04.06.2013 Views

MOE and MOR assessment technologies for improving graded ...

MOE and MOR assessment technologies for improving graded ...

MOE and MOR assessment technologies for improving graded ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

PRODUCTS & PROCESSING<br />

PROJECT NUMBER: PNB040-0708 NOVEMBER 2009<br />

<strong>MOE</strong> <strong>and</strong> <strong>MOR</strong> <strong>assessment</strong><br />

<strong>technologies</strong> <strong>for</strong> <strong>improving</strong><br />

<strong>graded</strong> recovery of exotic<br />

pines in Australia<br />

This report can also be viewed on the FWPA website<br />

www.fwpa.com.au<br />

FWPA Level 4, 10-16 Queen Street,<br />

Melbourne VIC 3000, Australia<br />

T +61 (0)3 9927 3200 F +61 (0)3 9927 3288<br />

E info@fwpa.com.au W www.fwpa.com.au<br />

`


<strong>MOE</strong> <strong>and</strong> <strong>MOR</strong> <strong>assessment</strong> <strong>technologies</strong> <strong>for</strong><br />

<strong>improving</strong> <strong>graded</strong> recovery of exotic pines in<br />

Australia<br />

Prepared <strong>for</strong><br />

Forest & Wood Products Australia<br />

by<br />

H. Bailleres, G. Hopewell <strong>and</strong> G. Boughton


Publication: <strong>MOE</strong> <strong>and</strong> <strong>MOR</strong> <strong>assessment</strong> <strong>technologies</strong> <strong>for</strong><br />

<strong>improving</strong> <strong>graded</strong> recovery of exotic pines in Australia<br />

Project No: PNB040-0708<br />

© 2009 Forest & Wood Products Australia Limited. All rights reserved.<br />

Forest & Wood Products Australia Limited (FWPA) makes no warranties or assurances with<br />

respect to this publication including merchantability, fitness <strong>for</strong> purpose or otherwise. FWPA <strong>and</strong><br />

all persons associated with it exclude all liability (including liability <strong>for</strong> negligence) in relation to<br />

any opinion, advice or in<strong>for</strong>mation contained in this publication or <strong>for</strong> any consequences arising<br />

from the use of such opinion, advice or in<strong>for</strong>mation.<br />

This work is copyright <strong>and</strong> protected under the Copyright Act 1968 (Cth). All material except the<br />

FWPA logo may be reproduced in whole or in part, provided that it is not sold or used <strong>for</strong><br />

commercial benefit <strong>and</strong> its source (Forest & Wood Products Australia Limited) is acknowledged.<br />

Reproduction or copying <strong>for</strong> other purposes, which is strictly reserved only <strong>for</strong> the owner or<br />

licensee of copyright under the Copyright Act, is prohibited without the prior written consent of<br />

Forest & Wood Products Australia Limited.<br />

ISBN: 978-1-920883-89-8<br />

Researcher:<br />

H. Bailleres, G. Hopewell<br />

Queensl<strong>and</strong> Primary Industries <strong>and</strong> Fisheries<br />

Gate 3/ 80 Meiers Rd<br />

Indooroopilly Q 4068<br />

Subcontractors:<br />

G. Boughton<br />

TimberED Services Pty Ltd<br />

PO Box 30<br />

Duncraig East WA 6023<br />

L. Branchiau,<br />

CIRAD Xylometry<br />

73, Rue J.F. Breton, BP 5035<br />

34032 Montpellier France<br />

Final report received by FWPA in November, 2009<br />

Forest & Wood Products Australia Limited<br />

Level 4, 10-16 Queen St, Melbourne, Victoria, 3000<br />

T +61 3 9927 3200 F +61 3 9927 3288<br />

E info@fwpa.com.au<br />

W www.fwpa.com.au


Executive Summary<br />

This project was designed to provide the structural softwood processing industry with the<br />

basis <strong>for</strong> improved green <strong>and</strong> dry grading to allow maximise MGP grade yields, consistent<br />

product per<strong>for</strong>mance <strong>and</strong> reduced processing costs. To achieve this, advanced statistical<br />

techniques were used in conjunction with state-of-the-art property measurement systems.<br />

Specifically, the project aimed to make two significant steps <strong>for</strong>ward <strong>for</strong> the Australian<br />

structural softwood industry:<br />

• <strong>assessment</strong> of <strong>technologies</strong>, both existing <strong>and</strong> novel, that may lead to selection of a<br />

consistent, reliable <strong>and</strong> accurate device <strong>for</strong> the log yard <strong>and</strong> green mill. The purpose is<br />

to more accurately identify <strong>and</strong> reject material that will not make a minimum grade of<br />

MGP10 downstream;<br />

• improved correlation of grading <strong>MOE</strong> <strong>and</strong> <strong>MOR</strong> parameters in the dry mill using new<br />

analytical methods <strong>and</strong> a combination of devices.<br />

The three populations tested were stiffness-limited radiata pine, strength-limited radiata pine<br />

<strong>and</strong> Caribbean pine. Resonance tests were conducted on logs prior to sawmilling, <strong>and</strong> on<br />

boards. Raw data from existing in-line systems were captured <strong>for</strong> the green <strong>and</strong> dry boards.<br />

The dataset was analysed using classical <strong>and</strong> advanced statistical tools to provide correlations<br />

between data sets <strong>and</strong> to develop efficient strength <strong>and</strong> stiffness prediction equations.<br />

Stiffness <strong>and</strong> strength prediction algorithms were developed from raw <strong>and</strong> combined<br />

parameters.<br />

Parameters were analysed <strong>for</strong> comparison of prediction capabilities using in-line parameters,<br />

off-line parameters <strong>and</strong> a combination of in-line <strong>and</strong> off-line parameters.<br />

The results show that acoustic resonance techniques have potential <strong>for</strong> log <strong>assessment</strong>, to sort<br />

<strong>for</strong> low stiffness <strong>and</strong>/or low strength, depending on the resource. From the log measurements,<br />

a strong correlation was found between the average static <strong>MOE</strong> of the dried boards within a<br />

log <strong>and</strong> the predicted value. These results have application in segregating logs into structural<br />

<strong>and</strong> non-structural uses. Some commercial <strong>technologies</strong> are already available <strong>for</strong> this<br />

application such as Hitman LG640.<br />

For green boards it was found that in-line <strong>and</strong> laboratory acoustic devices can provide a good<br />

prediction of dry static <strong>MOE</strong> <strong>and</strong> moderate prediction <strong>for</strong> <strong>MOR</strong>.There is high potential <strong>for</strong><br />

segregating boards at this stage of processing. Grading after the log breakdown can improve<br />

significantly the effectiveness of the mill. Subsequently, reductions in non-structural volumes<br />

can be achieved. Depending on the resource it can be expected that a 5 to 8 % reduction in<br />

non structural boards won’t be dried with an associated saving of $70 to 85/m 3<br />

For dry boards, vibration <strong>and</strong> a st<strong>and</strong>ard Metriguard CLT/HCLT provided a similar level of<br />

prediction on stiffness limited resource. However, Metriguard provides a better strength<br />

prediction in strength limited resources (due to this equipment’s ability to measure local<br />

characteristics). The combination of grading equipment specifically <strong>for</strong> stiffness related<br />

predictors (Metriguard or vibration) with defect detection systems (optical or X-ray scanner)<br />

provides a higher level of prediction, especially <strong>for</strong> <strong>MOR</strong>. Several commercial <strong>technologies</strong><br />

are already available <strong>for</strong> acoustic grading on board such those from Microtec, Luxscan,<br />

Falcon engineering or Dynalyse AB <strong>for</strong> example.<br />

i


Differing combinations of equipment, <strong>and</strong> their strategic location within the processing chain,<br />

can dramatically improve the efficiency of the mill, the level of which will vary depending of<br />

the resource. For example, an initial acoustic sorting on green boards combined with an<br />

optical scanner associated with an acoustic system <strong>for</strong> grading dry board can result in a large<br />

reduction of the proportion of low value low non-structural produced.<br />

The application of classical MLR on several predictors proved to be effective, in particular <strong>for</strong><br />

<strong>MOR</strong> predictions. However, the usage of a modern statistics approach (chemometrics tools)<br />

such as PLS proved to be more efficient <strong>for</strong> <strong>improving</strong> the level of prediction.<br />

Compared to existing <strong>technologies</strong>, the results of the project indicate a good improvement<br />

potential <strong>for</strong> grading in the green mill, ahead of kiln drying <strong>and</strong> subsequent cost-adding<br />

processes. The next stage is the development <strong>and</strong> refinement of systems <strong>for</strong> this purpose.<br />

ii


Table of Contents<br />

Executive Summary .................................................................................................................... i<br />

Introduction ................................................................................................................................1<br />

Literature review ........................................................................................................................ 3<br />

Radiata pine................................................................................................................................3<br />

Caribbean pine............................................................................................................................ 4<br />

Non-destructive testing methods................................................................................................ 4<br />

Radiography ........................................................................................................................... 5<br />

Evaluation of visual characteristics........................................................................................ 6<br />

Near-infrared (NIR) methods................................................................................................. 6<br />

Microwave techniques............................................................................................................ 6<br />

Stress wave methods .............................................................................................................. 7<br />

Ultrasonic ........................................................................................................................... 7<br />

Sonic................................................................................................................................... 8<br />

Resonance method.............................................................................................................. 8<br />

Remark on the measured acoustic velocities.................................................................... 10<br />

Machine stress rating............................................................................................................ 10<br />

Static bending........................................................................................................................... 11<br />

NDT of wood <strong>for</strong> grading: results from previous studies .................................................... 12<br />

St<strong>and</strong>ard requirements.............................................................................................................. 15<br />

Study Methodology.................................................................................................................. 16<br />

Materials............................................................................................................................... 16<br />

Sampling method.............................................................................................................. 16<br />

Kiln drying ....................................................................................................................... 18<br />

Equipment <strong>and</strong> methods....................................................................................................... 18<br />

Non destructive testing- mechanical properties measurement......................................... 18<br />

Machine stress rating............................................................................................................... 18<br />

In-line acoustic test .................................................................................................................. 19<br />

Off-line acoustic test................................................................................................................. 19<br />

Gamma ray............................................................................................................................... 22<br />

Non destructive testing- structural properties measurement............................................ 23<br />

Linear high grader (LHG)........................................................................................................ 23<br />

WoodEye ® ........................................................................................................................ 23<br />

Destructive st<strong>and</strong>ard static bending tests ......................................................................... 24<br />

Results ...................................................................................................................................... 29<br />

Static bending....................................................................................................................... 29<br />

Biased <strong>and</strong> r<strong>and</strong>om tests together..................................................................................... 29<br />

R<strong>and</strong>om <strong>and</strong> biased .......................................................................................................... 32<br />

Biased vs R<strong>and</strong>om............................................................................................................ 37<br />

Position of the boards within the logs .............................................................................. 39<br />

Individual NDT predictors <strong>for</strong> static <strong>MOE</strong> <strong>and</strong> <strong>MOR</strong> biased evaluation ............................ 40<br />

At dry mill level ............................................................................................................... 40<br />

Radiata E resource........................................................................................................... 40<br />

Radiata R resource........................................................................................................... 43<br />

Caribbean resource.......................................................................................................... 47<br />

At green mill level............................................................................................................ 50<br />

Radiata E resource........................................................................................................... 50<br />

Radiata R resource........................................................................................................... 52<br />

Caribbean resource.......................................................................................................... 53<br />

Combinations of NDT predictors <strong>for</strong> static <strong>MOE</strong> <strong>and</strong> <strong>MOR</strong> evaluation ............................. 55


Radiata E resource............................................................................................................ 55<br />

Radiata R resource ........................................................................................................... 57<br />

Caribbean resource........................................................................................................... 58<br />

Prediction of static bending <strong>MOE</strong> <strong>and</strong> <strong>MOR</strong> from logs....................................................... 60<br />

Radiata E .......................................................................................................................... 60<br />

Radiata R.......................................................................................................................... 60<br />

Caribbean ......................................................................................................................... 61<br />

Practical grading into MGP grades ...................................................................................... 62<br />

Using best vibration prediction <strong>for</strong> logs........................................................................... 62<br />

Using best prediction <strong>for</strong> dry boards NDT <strong>technologies</strong>...................................................... 63<br />

Discussion <strong>and</strong> conclusions...................................................................................................... 72<br />

Recommendations .................................................................................................................... 74<br />

References ................................................................................................................................ 75<br />

Internet references................................................................................................................ 77<br />

St<strong>and</strong>ards .............................................................................................................................. 77<br />

Personal communication ...................................................................................................... 78<br />

Acknowledgements .................................................................................................................. 79<br />

Researcher’s Disclaimer........................................................................................................... 80<br />

Appendix 1. Resonance method extracted signal descriptors.................................................. 81<br />

Appendix 2. WoodEye profile descriptions. ............................................................................ 83<br />

Appendix 3. Biased <strong>and</strong> r<strong>and</strong>om testing protocol. ................................................................... 84<br />

Radiata E resource................................................................................................................ 84<br />

Radiata R <strong>and</strong> Caribbean resources...................................................................................... 84


Introduction<br />

In Australia, a range of native <strong>and</strong> exotic softwood <strong>for</strong>ests <strong>and</strong> plantations provide an<br />

important source of fibre <strong>for</strong> sawn, round <strong>and</strong> panel products. The introduced Pinus species<br />

are grown in plantations managed specifically to produce high volumes of structural products<br />

primarily <strong>for</strong> the domestic construction sector. The native softwoods such as white cypress<br />

(Callitris glaucophylla Thompson & Johnson) <strong>and</strong> hoop pine (Araucaria cunninghamii Aiton<br />

ex D. Don) were not included in this study <strong>and</strong> all further reference to the term ‘softwood/s’<br />

in this document, indicates plantation-grown exotic pine.<br />

The scope of work was identified as a priority by the Australian softwood sector, <strong>and</strong> is the<br />

first study of its type conducted in Australia. Subsets of the radiata pine <strong>and</strong> Caribbean pine<br />

populations represent the range of mechanical properties of the majority of the Australiangrown<br />

exotic pine resource. This sector requires a universally applicable method to accurately<br />

<strong>and</strong> rapidly predict the strength <strong>and</strong> stiffness of green logs <strong>and</strong> boards processed <strong>for</strong> structural<br />

products. This will allow non-structural <strong>and</strong> low stiffness boards to be diverted be<strong>for</strong>e<br />

entering the dry chain.<br />

There is often a high variability in wood properties of fast grown trees harvested at a young<br />

age, even within a single stem (Zobel <strong>and</strong> Buijtenen, 1989). Consequently, within a log or<br />

even within a board, the wood properties may be significantly different. Wood can be<br />

characterized as a highly heterogeneous <strong>and</strong> highly anisotropic material. Heterogeneous<br />

means that wood does not have a uni<strong>for</strong>m structure <strong>and</strong> this variability can affect strength, <strong>for</strong><br />

example knots, resin pockets <strong>and</strong> reaction wood. Anisotropic means that wood is a very<br />

oriented material, in other words directionally dependent with different properties in different<br />

planes. The strength <strong>and</strong> stiffness in the longitudinal direction of the tree are much higher than<br />

in the transverse directions. This effect can cause problems when the grain direction is not<br />

always parallel to the sawn direction of boards. High slope of grain can seriously decrease the<br />

bending strength. Because such variation is not acceptable in wood used <strong>for</strong> structural<br />

applications, it must be appropriately <strong>graded</strong> to ensure safety <strong>and</strong> per<strong>for</strong>mance in service.<br />

Grading is the process by which timber is sorted into appropriate stress grades with consistent<br />

properties in each grade. Inevitably, there is a range of properties within a grade <strong>and</strong><br />

significant overlap in properties between groups. Currently, processors undertake the grading<br />

task in the dry mill. To prevent unnecessary processing, associated costs <strong>and</strong> energy usage,<br />

<strong>for</strong> example kiln drying of low strength <strong>and</strong> ultimately non-structural boards, the grading<br />

process <strong>and</strong> subsequent segregation should be done as early as possible in the value chain.<br />

This project was initiated <strong>and</strong> designed in order to address this issue.<br />

The main objectives of this project were:<br />

• To improve the use of robust predictors of strength <strong>and</strong> stiffness acquired through<br />

existing in-line processing equipment such as Metriguard Continuous Lumber Tester<br />

(CLT), knot area ratio (KAR), acoustics, gamma ray <strong>and</strong> optical scanners.<br />

• To improve the use of grading tools <strong>for</strong> upstream sorting: from logs, green boards <strong>and</strong><br />

finally dry mill products.<br />

• To identify if new parameters are available to refine current predictive mechanisms,<br />

<strong>and</strong> determine the most effective way to input these into prediction equations.<br />

• To define the accuracy of the relevant parameters, or combination of parameters, <strong>for</strong> a<br />

number of <strong>technologies</strong> to independently improve grading systems.<br />

1


• To develop advanced vibrational methods in order to grade softwood boards <strong>and</strong> logs<br />

<strong>for</strong> structural purposes.<br />

The trials discussed here included three separate samples from distinct populations: stiffness<br />

limited radiata pine, strength-limited radiata pine <strong>and</strong> Caribbean pine. For each sample, logs<br />

were measured <strong>and</strong> weighed to provide density data, then tested <strong>for</strong> acoustic resonance. The<br />

resulting <strong>MOE</strong> calculations were later correlated with results from reference tests on the<br />

boards sawn from the logs.<br />

After sawing, the green boards were subjected to a range of tests to provide relevant data as<br />

summarised below:<br />

• Industrial acoustic test (Weyerhaeuser Thumper <strong>for</strong> <strong>MOE</strong>);<br />

• Gamma ray (Geological & Nuclear sciences Ltd <strong>for</strong> moisture content <strong>and</strong> density);<br />

• Metriguard Continuous Lumber Tester (<strong>MOE</strong>);<br />

• Metriguard High Capacity Lumber Tester (<strong>MOE</strong>);<br />

• LHG X-ray (<strong>for</strong> density <strong>and</strong> knot area ratio, strength-limited radiata pine only);<br />

• WoodEye ® (laser <strong>and</strong> camera optical scanner <strong>for</strong> defect type, size <strong>and</strong> position at<br />

production speed);<br />

• Acoustic measurements (Bing ® ; longitudinal <strong>and</strong> transverse <strong>MOE</strong>s).<br />

• Reference tests (<strong>MOE</strong> <strong>and</strong> <strong>MOR</strong>, AS 4063:1992).<br />

Reference tests (static bending <strong>for</strong> <strong>MOE</strong> <strong>and</strong> <strong>MOR</strong>, AS 4063:1992) including both biased <strong>and</strong><br />

r<strong>and</strong>om samples, were undertaken on a universal static 4-point bending test machine <strong>and</strong><br />

visual defects were measured.<br />

The dataset was analysed using mathematical <strong>and</strong> chemometrics’ statistical tools to extract<br />

relevant parameters <strong>and</strong> to provide correlations between data sets <strong>and</strong> develop efficient<br />

strength <strong>and</strong> stiffness prediction equations. Stiffness <strong>and</strong> strength prediction algorithms were<br />

developed from raw <strong>and</strong> combined parameters using specific chemometrics’ tools including<br />

multi-variate linear regression (MLR) <strong>and</strong> partial least squares (PLS).<br />

The parameters were analysed <strong>for</strong> comparison of prediction capabilities from in-line<br />

parameters, off-line parameters (vibration analysis <strong>and</strong> manual defects measurements) <strong>and</strong> a<br />

combination of in-line <strong>and</strong> off-line parameters.<br />

The predictors which provided the best correlations to the static <strong>MOE</strong> were used to sort the<br />

boards into MGP grades<br />

The results from the project will allow processors to improve existing machine grading <strong>and</strong><br />

assist in the development of next generation systems <strong>for</strong> more accurate grading of structural<br />

wood. The improved confidence in estimating strength <strong>and</strong> stiffness values <strong>for</strong> dried timber<br />

will allow <strong>for</strong> grading closer to threshold limits, thus <strong>improving</strong> structural grade yields<br />

resulting in a more efficient <strong>and</strong> profitable use of the softwood resource. This equates to<br />

greater resource optimisation, reduced costs <strong>and</strong> increased profits <strong>for</strong> the softwood sector.<br />

2


Literature review<br />

Radiata pine<br />

Pinus radiata D. Don is marketed in Australia as radiata pine , but it is also known as<br />

Monterey pine <strong>and</strong> Insignis pine. It is easy to establish, can grow under a range of site<br />

conditions <strong>and</strong> produces large quantities of useable wood over relatively short rotations.<br />

Radiata pine from two disjunct sources viz south east Australia <strong>and</strong> south-west Australia, was<br />

investigated. Radiata pine is a versatile <strong>and</strong> widely planted species <strong>and</strong> in Australia there are<br />

almost 720,000 hectares (Web 1, 2003) of established radiata pine plantation <strong>for</strong>ests.<br />

Native radiata pine occurs at five locations in central to north America. Three of these<br />

locations are in Cali<strong>for</strong>nia: Año Nuevo, the Monterey Peninsula <strong>and</strong> Cambria. The other two<br />

are found on the two small isl<strong>and</strong>s off the coast of Mexico, Cedros (Pinus radiata var.<br />

cedrosensis) <strong>and</strong> Guadalupu (Pinus radiata var. binata) (Bootle, 2005).<br />

The wood of radiata pine is pale yellow-brown <strong>and</strong> is generally straight grained with<br />

prominent growth rings <strong>for</strong>med by alternating b<strong>and</strong>s of earlywood <strong>and</strong> latewood. It has low<br />

natural durability, however may be treated with preservatives <strong>for</strong> outdoor use. The wood has a<br />

good strength to weight ratio, with good nail holding <strong>and</strong> gluing ability, resistance to nail<br />

splitting <strong>and</strong> is relatively easy to saw <strong>and</strong> dry (Bootle, 2005).<br />

The air-dry density (12% moisture content) is variable but typically is around 545 kg/m 3 <strong>and</strong><br />

nominated strength groups are S6, SD6 (Hopewell, 2006). Wood density increases as the tree<br />

ages, <strong>and</strong> is influenced by environmental factors. New Zeal<strong>and</strong> research has found that trees<br />

grown at lower altitude in warmer areas have a higher wood density than those grown at<br />

higher altitudes in cooler areas (Kininmonth <strong>and</strong> Whitehouse, 1991).<br />

The innermost annual rings in the tree stem have a different anatomical structure compared to<br />

the wood in outer layers of the stem. The wood in the innermost rings is known as ‘juvenile<br />

wood’ <strong>and</strong> it has significantly different mechanical properties than the outer ‘adult wood’.<br />

The central core generally exhibits pronounced spiral grain, shorter fibres <strong>and</strong> lower wood<br />

density (as low as 350 kg/m 3 at 12% MC). This corewood is usually confined to the first ten<br />

to twenty growth rings only <strong>and</strong> is regarded as low-quality wood with the following<br />

characteristics (Ilic et al, 2003, except as noted):<br />

• wide growth-rings;<br />

• high grain spirality;<br />

• low density <strong>and</strong> stiffness;<br />

• thin cell walls <strong>and</strong> short trachieds;<br />

• high longitudinal shrinkage;<br />

• low transversal shrinkage<br />

• presence of compression wood, <strong>and</strong><br />

• lower cellulose:lignin ratio (Bendtsen, 1978).<br />

Radiata pine dries rapidly, <strong>and</strong> is usually kiln dried from the green condition at high<br />

temperatures e.g. 140°C. The wood is easy to dry but boards sawn from the central core zone<br />

which are prone to distortion. Improved stability of the seasoned product is achieved by presteaming<br />

<strong>for</strong> several hours <strong>and</strong> the use of concrete stack weights during drying (Bootle, 2005).<br />

Radiata pine has a wide range of structural <strong>and</strong> decorative uses including framing, furniture,<br />

paneling, lining, glued laminated beams, veneer, plywood <strong>and</strong> pulp. When treated with<br />

3


preservatives, it can be used <strong>for</strong> outdoor applications such as posts <strong>and</strong> poles (Bootle, 2005).<br />

Small diameter logs produced from thinning operations can be used <strong>for</strong> posts or pulpwood.<br />

Caribbean pine<br />

Caribbean pine sourced from south-east Queensl<strong>and</strong> provided the balance of test material <strong>for</strong><br />

the trials. Caribbean pine has been planted in Queensl<strong>and</strong> <strong>and</strong> New South Wales where it has<br />

developed a reputation <strong>for</strong> excellent growth with minimal branching, even on poorly drained<br />

soils, resulting in desirable wood quality viz smaller <strong>and</strong> less frequent knots than related Pinus<br />

species.<br />

Caribbean pine is pale yellow to yellow <strong>for</strong> the sapwood, <strong>and</strong> yellow to pale brown, with pink<br />

tints <strong>for</strong> the heartwood. The difference of colour between earlywood <strong>and</strong> latewood is<br />

pronounced, <strong>and</strong> the grain is straight with a coarse <strong>and</strong> uneven texture.<br />

It has a low natural durability but can be effectively impregnated with chemical preservatives.<br />

The air-dry density ranges from 545 to 575 kg/m³ (Hopewell, 2006). Provisional strength<br />

groups have been assigned to Caribbean pine as (S6) <strong>and</strong> (SD6).<br />

Caribbean pine dries rapidly, <strong>and</strong> is usually kiln dried from the green condition at high to very<br />

high temperatures e.g.140-180°C without loss of strength (Siemon, 1981). The wood is easy<br />

to dry, except boards sawn from the central core zone which are prone to distortion. Improved<br />

stability of the seasoned product is achieved by reconditioning <strong>and</strong> the use of stack weights<br />

during drying.<br />

Caribbean pine has a wide range of uses including framing, flooring, mouldings, joinery,<br />

furniture, plywood, treated l<strong>and</strong>scaping <strong>and</strong> roundwood products, laminated beams, medium<br />

density fibreboard <strong>and</strong> paper production. Resin can be a problem during sawmilling as it<br />

builds up on saws <strong>and</strong> other processing equipment (Bootle, 2005).<br />

Non-destructive testing methods<br />

Non-destructive testing (NDT), also called non-destructive evaluation (NDE) <strong>and</strong> nondestructive<br />

inspection (NDI), is the science of identifying physical <strong>and</strong> mechanical properties<br />

of a piece of material without altering its end-use capabilities (Ross <strong>and</strong> Pellerin, 1994).<br />

Since the 1920’s, NDT methods have developed from laboratory testing to an indispensable<br />

production tool. Often components are too costly, or destructive testing is not possible, thus<br />

NDT is becoming increasingly important as a quality control management tool in almost<br />

every manufacturing process.<br />

Today there are a large variety of NDT methods which are used worldwide to detect material<br />

characteristics such as: variation in structure, the presence of cracks or other physical or<br />

mechanical discontinuities, dimensions of products <strong>and</strong> to determine other characteristics of<br />

material. The most common methods are listed below (Bucur, 2003):<br />

• visual <strong>and</strong> optical testing;<br />

• ultrasonic testing;<br />

• radiographic testing;<br />

• impulse excitation technique;<br />

• electromagnetic testing;<br />

• vibration methods;<br />

4


• magnetic resonance;<br />

• laser testing;<br />

• near-infrared.<br />

Many of these methods originate from the medical field <strong>and</strong> have been adapted <strong>for</strong> industrial<br />

use (<strong>for</strong> example computed radiography <strong>and</strong> ultrasonic). These <strong>and</strong> other techniques are used<br />

in a wide range of industrial activities including aviation, aerospace, construction,<br />

manufacturing, pipelines <strong>and</strong> railways. Each method has advantages <strong>and</strong> disadvantages, <strong>for</strong><br />

example radiographic methods can be used on several materials to detect the internal<br />

condition <strong>and</strong>/or properties of samples but is limited by the thickness of the test material <strong>and</strong><br />

requires high safety precautions. Some methods are more suitable <strong>for</strong> detection <strong>and</strong> evaluation<br />

of certain flaws/properties than others. There<strong>for</strong>e, the right choice or the combination of<br />

methods is desirable.<br />

Using NDT to evaluate the physical properties of wood has its’ origin in the need to solve<br />

practical problems without destruction of the integrity of the object under inspection (Bucur,<br />

2003). The heterogeneity <strong>and</strong> anisotropy of wood make it difficult <strong>for</strong> manufacturers of <strong>for</strong>est<br />

products to provide a consistent quality to their customers. Many of the methods listed above<br />

have been adapted <strong>for</strong> predicting the per<strong>for</strong>mance of wood, but its wide variability makes it<br />

more challenging than <strong>for</strong> homogenous materials like metal <strong>and</strong> plastics (Ross <strong>and</strong> Pellerin,<br />

1994). There<strong>for</strong>e, research work on NDT of wood is required to determine a more accurate<br />

per<strong>for</strong>mance of a wood member (Ibid). Among the wood characteristics to be assessed nondestructively,<br />

strength <strong>and</strong> stiffness are the most important <strong>for</strong> structural applications. The<br />

only way to determine the true strength of a piece of timber is to break it. But afterwards it is<br />

of no use as a load carrying component. There<strong>for</strong>e predicting the material characteristics of<br />

wood through NDT techniques is vital <strong>for</strong> the timber industry <strong>and</strong> has a long history of<br />

application in the wood products industry (Halabe et al, 1995). Moreover, the use of structural<br />

components is generally under st<strong>and</strong>ard applications which define the per<strong>for</strong>mance of the<br />

product in regard to its purpose.<br />

It has been shown that density, knots (size, frequency, <strong>and</strong> location) <strong>and</strong> modulus of elasticity<br />

(<strong>MOE</strong>) are the most suitable parameters <strong>for</strong> wood strength prediction. The modulus of<br />

elasticity has shown the best correlation to the strength <strong>for</strong> a single parameter (Steiger, 1996).<br />

Today a wide variety of NDT methods <strong>for</strong> wood are known <strong>and</strong> available on the market<br />

including:<br />

• radiography (X-ray <strong>and</strong> gamma-ray);<br />

• evaluation of visual characteristics;<br />

• near-infrared;<br />

• microwave;<br />

• ultrasonic;<br />

• acoustic;<br />

• machine stress rating (MSR);<br />

• reference testing (quasi-static test).<br />

Further <strong>technologies</strong> exist which are either not yet fully developed, or currently too complex<br />

or expensive to be used in commercial applications.<br />

Radiography<br />

In the 1980’s medical X-ray <strong>and</strong> gamma ray scanners were first used on wood. Scientists soon<br />

realised that the physical properties of wood are similar to the human body <strong>and</strong> thus<br />

developed a non-medical scanner which could be used in the timber industry. X-ray <strong>and</strong><br />

5


gamma ray scanning identifies internal heterogeneities <strong>and</strong> defects in logs <strong>and</strong> sawn boards<br />

(Bucur, 2003). The rays are able to penetrate material without being reflected or broken. On<br />

the basis of this feature a precise map of the internal heterogeneities is created. Depending on<br />

the system it is possible to acquire a 2D or 3D image. Such industrial wood scanner systems<br />

require a high X-ray or gamma ray source.<br />

Gamma rays are of lower intensity than X-rays <strong>and</strong> can be used as a portable system <strong>for</strong><br />

inspection of trees, poles <strong>and</strong> building elements (Bucur, 2003). The scan is done by irradiating<br />

the specimen from one or more sides <strong>and</strong> collecting the intensity of the transmitted radiation<br />

on the opposite side. The absorbed radiation is linked with the density <strong>and</strong> the moisture<br />

content of the material (Duff, 2005). Using this technology several strength affecting<br />

parameters like knots, density <strong>and</strong> moisture content can be determined contact free. The main<br />

advantages are that data are available in real time <strong>and</strong> a large volume of material can be<br />

rapidly inspected (Hanhijärvi et al, 2005). The disadvantage is that to ensure safety, the high<br />

intensity equipment has relatively large space requirements (Bucur, 2003).<br />

Evaluation of visual characteristics<br />

Visual inspection is one of the simplest <strong>and</strong> oldest methods of detecting exterior defects in<br />

wood members. The inspector observes the sample <strong>for</strong> parameters affecting strength such as<br />

knots, slope of grain <strong>and</strong> decay. This method requires good lighting, close attention from<br />

experienced operators <strong>and</strong> is limited to external degrade <strong>and</strong> features (Ross et al, 2006).<br />

Visual inspection is a relatively subjective method.<br />

Automatic optical scanning systems allow this process at production speeds. Normally such<br />

machines are equipped with different camera <strong>and</strong> laser systems. Such cameras are also used to<br />

identify the variation in surface colour along the boards (Duff, 2005). These systems both<br />

detect the existence of the defect <strong>and</strong> its position. The data is usually sent directly to a<br />

docking machine to optimise the cutting process.<br />

Near-infrared (NIR) methods<br />

The spectroscopy of near-infrared waves ranges from 800 to 2500 nm. The advantage is that<br />

NIR can penetrate deeper than mid-infrared radiation <strong>and</strong> is simple to operate (minimal<br />

preparation <strong>and</strong> rapid measurement). NIR has mainly been used <strong>for</strong> non-destructive testing of<br />

organic materials such as agricultural or food products. Today it is used mainly in the medical<br />

<strong>and</strong> chemical fields, but recently NDT methods using NIR have been tested in the timber<br />

industry.<br />

Tsuchikawa (2006) presented recent technical <strong>and</strong> scientific reports of NIR spectroscopy<br />

research in the wood <strong>and</strong> paper science industries. Others described a near-infrared method<br />

developed to predict the lignin content of solid wood. Strong correlations were found between<br />

the predicted lignin contents <strong>and</strong> the contents obtained from a traditional chemical method<br />

(Yeh et al, 2004). Schimleck et al (2002) described NIR measurements on small clear<br />

samples of Eucalyptus delegatensis <strong>and</strong> Pinus radiata to determine a number of physical<br />

properties including density, <strong>MOE</strong>, micro fibril angle <strong>and</strong> modulus of rupture (<strong>MOR</strong>). Good<br />

correlations were observed <strong>for</strong> all parameters, with R 2 ranging from 0.77 <strong>for</strong> <strong>MOR</strong> through<br />

0.90 <strong>for</strong> <strong>MOE</strong> to 0.93 <strong>for</strong> density.<br />

Microwave techniques<br />

Similar to X-ray <strong>and</strong> gamma ray, microwaves penetrate the wood <strong>and</strong> on the basis of wave<br />

reflection, <strong>and</strong> checks, splits <strong>and</strong> irregularities in the test specimen can be detected. Compared<br />

to X-ray <strong>and</strong> gamma ray, this technology is less expensive, but at current production speeds<br />

6


the NDT of timber by microwaves is impeded/disturbed by mechanical vibrations (Bucur,<br />

2003). Microwave testing is a contact free method that can detect defects <strong>and</strong> irregularity<br />

within the wood based on the determination of its dielectric properties which are principally<br />

due to the water content. Baradit et al (2006) described a microwave technique to generate<br />

<strong>and</strong> process data on knots in wood be<strong>for</strong>e processing.<br />

Stress wave methods<br />

The analysis of mechanical wave propagation in media of various complexities enables the<br />

measurement of elastic properties (Bucur, 2003). The use of acoustic methods, vibrational in<br />

the audible range (sonic) <strong>and</strong> at a frequency beyond human hearing capability (ultrasonic), <strong>for</strong><br />

characterisation of the mechanical behavior of solid wood <strong>and</strong> wood-based composites is well<br />

documented. The velocity at which a stress wave travels in a member is dependent upon the<br />

properties of the member only. The terms sonic <strong>and</strong> ultrasonic refer only to the frequency of<br />

excitation used to introduce a wave into the member. All commercially available timing units,<br />

if calibrated <strong>and</strong> operated according to the manufacturer’s recommendations, yield to<br />

comparable results.<br />

The <strong>MOE</strong> <strong>and</strong> density are the main parameters which describe the wave propagation (Steiger,<br />

1996). The wave speed <strong>for</strong> isotropic <strong>and</strong> homogenous material at given physical conditions is<br />

defined by the following equation:<br />

Where,<br />

2<br />

2<br />

∂ u 1 ∂ u<br />

= 2<br />

2<br />

∂x<br />

V ∂t<br />

2<br />

x<br />

⇒<br />

u is longitudinal displacement<br />

t is the time measured<br />

Ex is the modulus of elasticity<br />

ρ is the wood density<br />

Vx is the wave propagation speed<br />

7<br />

V<br />

x<br />

=<br />

Ex<br />

ρ<br />

(Equation 1)<br />

Basically there are two different methods to exploit stress wave velocity measurement<br />

(Krautkrämer, 1996):<br />

• Transmission technique where the ultrasonic wave is introduced by a transmission<br />

head into the test piece <strong>and</strong> received by a second receiver head.<br />

• The pulse echo method where a wave is reflected in a salient place (defect, surface)<br />

<strong>and</strong> the echo by the combined sending/receipt head is registered.<br />

In the wood industry the transmission technique is more commonly used to indicate the<br />

stiffness of a sample (Steiger, 1996).<br />

Ultrasonic<br />

Ultrasonic describes a stress wave with a frequency greater than the upper limit of human<br />

hearing, that is, high frequency (20 kHz – 600 kHz <strong>and</strong> higher) stress wave. This technique is<br />

particularly useful to detect decay as the stress wave propagation is sensitive to the presence<br />

of degradation in wood. In general ultrasonic stress waves travel faster in high quality <strong>and</strong><br />

stiffness material than in material that is deteriorated or of low quality (Ross et al, 2006).<br />

Thus the wave propagation is based on physical <strong>and</strong> mechanical characteristics such as<br />

(Bucur, 2003):<br />

• <strong>MOE</strong>;


• density;<br />

• moisture content;<br />

• dimensions of the sample;<br />

• material type;<br />

• experimental conditions.<br />

Timber grading using ultrasound has been discussed by S<strong>and</strong>oz (1989) <strong>and</strong> in Steiger, (1991).<br />

It was found that by measuring spruce (Picea spp.) beams of 10 cm by 22 cm cross section<br />

<strong>and</strong> 4.40 m length, a correlation of R 2 = 0.46 between wave velocity <strong>and</strong> <strong>MOR</strong> existed.<br />

Hanhijärvi et al (2005) reported R 2 s of 0.42 <strong>for</strong> strength <strong>and</strong> 0.57 <strong>for</strong> stiffness using a<br />

Sylvamatic strength grading machine.<br />

Sonic<br />

Measurement of acoustic velocity using the stress wave method is based on the same principle<br />

as the ultrasonic method. In the acoustic domain the input consists of a low frequency<br />

(approximately 1 Hz – 20 kHz) stress wave.<br />

The wave is introduced to the material by striking the specimen with an impact hammer. A<br />

<strong>for</strong>ce transducer connected to the specimen records the input signal. On the other side of the<br />

specimen an accelerometer is connected which receives an output signal. The measurement of<br />

these two signals (input <strong>and</strong> output), allows the stress wave velocity to be calculated. The<br />

analysis is based on a time analysis, which doesn’t require the knowledge of the boundary<br />

(Brancheriau, 2007). When striking the specimen a range of stress wave frequencies, <strong>and</strong> thus<br />

velocities, is inducted. As a consequence several velocities can be measured.<br />

Most of the algorithms used extract the fastest speed <strong>and</strong> the average speed of the group. On<br />

the basis of the velocity <strong>and</strong> sample density the dynamic <strong>MOE</strong> can be assessed. The wave<br />

velocity is expressed as:<br />

L<br />

Vx = (Equation 2)<br />

t<br />

Where, Vx is the wave propagation speed<br />

L is the distance between the two probes (sensors)<br />

t is the time-of-flight (TOF)<br />

The wave velocity is linked with the <strong>MOE</strong> <strong>and</strong> density <strong>and</strong> can be calculated by Equation 1<br />

displayed above. This expression is exact only <strong>for</strong> isotropic <strong>and</strong> homogenous materials at<br />

given physical conditions <strong>and</strong> is there<strong>for</strong>e only approximate <strong>for</strong> wood. Further parameters like<br />

energy loss can also be extracted in order to access the non-homogeneity of the tested<br />

material. A couple of commercial equipments mainly develop <strong>for</strong> st<strong>and</strong>ing trees stress wave<br />

velocity measurement are available (Fakkop FRS-06/00 or Director ST300 <strong>for</strong> example).<br />

Resonance method<br />

The most convenient method <strong>for</strong> measuring <strong>MOE</strong> with high precision depends upon<br />

measurements of the resonance frequencies in different modes (longitudinal, flexion or<br />

torsional) of simple structures <strong>for</strong> which the geometry <strong>and</strong> boundary conditions are known<br />

(Brancheriau <strong>and</strong> Bailleres, 2002). The fact that the technique is based on resonant structure<br />

ensures that frequency measurements will be precise. The technique can be extended to<br />

measure damping parameters <strong>and</strong> several signal descriptors.<br />

8


To illustrate these methods consider a prismatic, homogenous <strong>and</strong> isotropic beam with a<br />

length L, height h <strong>and</strong> width w. After an impact hits the beam either longitudinally or<br />

laterally, it vibrates freely in compression or bending respectively. For a longitudinal wave<br />

(also known as compression or P-wave) the particles vibrate parallel to the direction the wave<br />

is travelling. In transversal wave (also known as shear wave or S-wave), the motion of the<br />

particles is perpendicular to its direction of propagation.<br />

Because of these different kinds of movements, the longitudinal method is used to estimate<br />

the compression <strong>and</strong> tension characteristics, while the transversal method determines the<br />

bending characteristics. By measuring the movement of a vibrating beam the fundamental<br />

resonant frequency can be determined by a Fast Fourier Trans<strong>for</strong>m algorithm. The following<br />

expression shows the relationship between frequency <strong>and</strong> speed:<br />

n<br />

n = V , n∈<br />

N*<br />

(Equation 3)<br />

2L<br />

f X<br />

Where, L is the length<br />

fn is the natural frequency (rank n)<br />

n is the frequency rank<br />

Vx is the wave propagation speed<br />

The dynamic modulus of elasticity along the longitudinal direction of the beam produced by a<br />

compression stress can be calculated using the following equation (Ibid):<br />

2<br />

2 fn<br />

E = 4L<br />

ρ<br />

(Equation 4)<br />

2<br />

n<br />

Where, E is the dynamic <strong>MOE</strong><br />

ρ is the wood density<br />

fn is the natural frequency (rank n)<br />

n is the number of frequencies<br />

Using the transversal measurement, produced by a flexion stress, we can apply Bernoulli’s<br />

model which provides a solution to calculate the dynamic modulus of elasticity. This is<br />

achieved using the following equation (Ibid):<br />

E<br />

Gz<br />

2<br />

n<br />

4<br />

2 ρAL<br />

f<br />

= 4π<br />

(Equation 5)<br />

I P<br />

Where, E is the dynamic modulus of elasticity<br />

ρ is the wood density<br />

A is cross-section area<br />

L is the length<br />

fn is the natural frequency (rank n)<br />

Pn is solution of Bernoulli (rank n)<br />

IGz is the moment of inertia, which can be calculated <strong>for</strong> a rectangular section<br />

as follows:<br />

9<br />

n


3<br />

bh<br />

IGz = (Equation 6)<br />

12<br />

Where, b is base, horizontal dimension (length)<br />

h is height, vertical dimension (length)<br />

12 is the constant <strong>for</strong> moment of inertia of a member with rectangular cross<br />

section.<br />

In Timoshenko’s model <strong>for</strong> transversal / flexion vibration, the shear effect is no longer<br />

ignored. The solution according to Bordonné is more accurate <strong>for</strong> higher depth to length ratio<br />

beams. It allows the extraction of the shear <strong>MOE</strong> <strong>and</strong> uses more resonance frequencies<br />

in<strong>for</strong>mation (Brancheriau <strong>and</strong> Bailleres, 2002).<br />

Recent studies (Ross et al, 2005; Tsehaye et al, 2000; Wang et al, 2003) have demonstrated<br />

that predicting <strong>MOE</strong> of trees, logs <strong>and</strong> sawn timber using acoustic velocities as well as<br />

resonance methods are highly correlated with the static <strong>MOE</strong> measured. Halabe et al (1995)<br />

found coefficients of determinations (R 2 ) between static bending <strong>MOE</strong> <strong>and</strong> stress wave <strong>MOE</strong><br />

of 0.73 <strong>and</strong> 0.74 in the green <strong>and</strong> dry stage, respectively.<br />

Several commercial <strong>technologies</strong> are already available <strong>for</strong> acoustic grading on board such<br />

those from Microtec, Luxscan, Falcon engineering or Dynalyse AB <strong>for</strong> example.<br />

Remark on the measured acoustic velocities<br />

From above there are two different methods to measure the stress wave velocity:<br />

1- By measuring the time-of-flight (TOF) according to equation 2. The accuracy of TOF<br />

measurement depends on accurate identification of the arrival times of the acoustic<br />

wave signals, each from a start sensor (impact hammer) <strong>and</strong> a stop sensor<br />

(accelerometer). It depends on the quality of the signal recorded <strong>and</strong> the extraction<br />

algorithm used to detect the start <strong>and</strong> the stop signals. The <strong>MOE</strong> can be calculated by<br />

applying equation 1 provided that the density is known. The TOF method applied on<br />

st<strong>and</strong>ing tree is likely disturbed by dilatational or quasi-dilatational waves rather than<br />

one-dimensional plane waves. This leads to st<strong>and</strong>ing tree velocity being significantly<br />

higher than log velocities <strong>and</strong> skewed relationship between tree <strong>and</strong> log acoustic<br />

measurements.<br />

2- By measuring the resonance frequency according to equation 3. The resonance-based<br />

acoustic method is a well-established NDT technique <strong>for</strong> measuring long, slender<br />

wood members. The inherent accuracy <strong>and</strong> robustness of this method provide a<br />

significant advantage over TOF measurement in applications such as log<br />

measurement. In contrast to TOF measurement, the resonance approach stimulates<br />

many, possibly hundreds, of acoustic pulse reverberations in a log, resulting in a very<br />

accurate <strong>and</strong> repeatable velocity measurement. Because of this accuracy, the acoustic<br />

velocity of the logs obtained by the resonance-based acoustic method is usually a<br />

better predictor than the velocity obtained by measuring the TOF. The <strong>MOE</strong> can be<br />

calculated by applying equation 4 provided that the density is known. The constraint<br />

linked to this method is that the boundary conditions have to be perfectly known, this<br />

is not possible on st<strong>and</strong>ing tree.<br />

Machine stress rating<br />

Static bending is the foundation of MSR of timber. The development of such machines started<br />

in the 1960s <strong>and</strong> by 1963 the first industrial MSR machines were operating in the USA. These<br />

10


machines were characterised by a very high efficiency, but they could only sort planed timber<br />

with a maximum depth of approximately 40 mm. There<strong>for</strong>e these machines found no practical<br />

application in Europe where larger dimensions are commonly produced <strong>and</strong> used (Krzosek,<br />

2003).<br />

MSR is currently the most common dynamic, mechanical load procedure. It measures the<br />

local stiffness of the material <strong>and</strong> sorts it into various <strong>MOE</strong> classes. The boards are fed<br />

through the machine flat wise, longitudinally <strong>and</strong> bent by rollers upwards <strong>and</strong> downwards in<br />

two sections. The span between the rollers is typically around 1.2 m.<br />

Depending on the design, the machine bends the boards to a constant deflection <strong>and</strong> measures<br />

the required <strong>for</strong>ce or the machine bends the boards with a constant <strong>for</strong>ce <strong>and</strong> measures the<br />

deflection. Using the load deflection relationship, the local <strong>MOE</strong> can be determined directly<br />

by using equations sourced from fundamental mechanics of material on every point of the<br />

board except <strong>for</strong> approximately the first <strong>and</strong> last 500 mm. This test method allows the<br />

stiffness profile of a board to be determined. Boards are usually <strong>graded</strong> using a combination<br />

of the lowest <strong>MOE</strong> (Lowpoint <strong>MOE</strong>), its position <strong>and</strong> the average <strong>MOE</strong> of the boards (Duff,<br />

2005).<br />

Static bending<br />

The resistance of a sample to slowly applied loads is measured by the static bending test. This<br />

procedure is generally conducted under st<strong>and</strong>ard conditions such as 20°C air temperature,<br />

65% relative air humidity <strong>and</strong> 12% wood moisture content. These values vary slightly<br />

depending on the st<strong>and</strong>ard used. The ends of the test sample are supported on rollers <strong>and</strong> a<br />

load is applied either centrally (three point bending) or two loads are applied in the middle<br />

third of the span (four point bending). In the past, three point bending tests were used to test<br />

small clear wood samples or wood composites, whereas four point bending tests were used to<br />

test full size specimens, although four point bending can also be used <strong>for</strong> small clear wood<br />

specimens.<br />

Because the values obtained by the different methods cannot be directly compared,<br />

Brancheriau et al (2002) presented an analytic <strong>for</strong>mula using a crossing coefficient between 3<br />

point <strong>and</strong> 4 point bending. The deflection <strong>and</strong> load are measured at intervals using a<br />

deflectometer <strong>and</strong> load cell respectively.<br />

The first part of the curve is a straight line. There the deflection is directly proportional to the<br />

load <strong>and</strong> once the load is removed, the test specimen will return to its original state. This<br />

de<strong>for</strong>mation is there<strong>for</strong>e in the elastic part of the beam (ε elastic). With increasing load a limit<br />

point of proportionality is reached (σp). Afterwards the de<strong>for</strong>mation is no longer proportional<br />

to the load. With further load increase the material becomes plastic. This means when the load<br />

is removed, the beam will not return to its initial state (ε plastic). By increasing the stress to<br />

the point of maximum load (σu), the material begins to yield <strong>and</strong> fracture. The static <strong>MOE</strong> is<br />

determined from the slope α <strong>and</strong> the <strong>MOR</strong> value is equivalent to the maximum load attained.<br />

These values are considered as reference values.<br />

The test span of such measurements depends on the dimensions of the specimen. Normally it<br />

is 18 times the depth of the sample. Thus, long thin specimens need to be docked to the<br />

required length. In the different st<strong>and</strong>ards, a variety of methods <strong>for</strong> selecting the test specimen<br />

from a piece of timber is defined (Leicester, 2004). For example in EN 408 a selection from<br />

the low point of the board is required (EN408, 1995).<br />

11


Low point is the location within the board with the lowest grading modulus also known as<br />

biased position test sampling. Australian New Zeal<strong>and</strong> St<strong>and</strong>ard AS/NZS4063 specifies that<br />

the sample is selected r<strong>and</strong>omly from a piece of timber (AS/NZS 4063, 1992). R<strong>and</strong>om<br />

sampling <strong>and</strong> r<strong>and</strong>om position testing gives the direct comparison with the design values.<br />

However, the sampling error <strong>for</strong> strength data collected this way can be relatively high<br />

(Boughton, pers. comm.).<br />

R<strong>and</strong>om position tests are useful to provide a good approximation to population average<br />

<strong>MOE</strong>. Biased position tests can focus on the low end of the distribution more effectively <strong>and</strong><br />

are useful to determine the effectiveness of grading systems (Leicester, 2004). However, the<br />

data from biased position tests cannot be compared directly with the Australian design values<br />

so acceptance criteria <strong>for</strong> these tests must be found from comparative testing on each size <strong>and</strong><br />

grade of product. These two methods may lead to considerably different results of mechanical<br />

properties. Leicester et al (1996) studied the equivalence between different in-grade testing<br />

methods. <strong>MOE</strong> <strong>and</strong> <strong>MOR</strong> were measured on about 150, 90 x 35 mm F5 <strong>and</strong> F8 grades of<br />

radiata pine boards with a length of 4.8 m. Small but significant differences of approximately<br />

20 percent at both the mean <strong>and</strong> the 5-percentile <strong>for</strong> strength were found.<br />

NDT of wood <strong>for</strong> grading: results from previous studies<br />

Brancheriau <strong>and</strong> Bailleres (2003) per<strong>for</strong>med dynamic tests on 96 structural boards of larch<br />

(Larix europeaea). Through Partial Least Squares (PLS) analysis of acoustic vibrations in the<br />

audible frequency range, the researchers showed that the stiffness <strong>and</strong> strength could be<br />

accurately estimated. Further, they suggested that rapid, in-line grading systems could be<br />

developed relatively inexpensively through the installation of a vibration sensor, acquisition<br />

card <strong>and</strong> a computer <strong>for</strong> Fast Fourier Trans<strong>for</strong>ms <strong>and</strong> matrices’ calculations.<br />

One hundred southern pine (Pinus spp.) boards of dimension 100 x 50 mm <strong>and</strong> a length of 2.4<br />

m were tested in green <strong>and</strong> dry states by Halabe et al (1995). The velocity of longitudinal<br />

stress waves <strong>and</strong> the dynamic <strong>MOE</strong> were measured with transverse vibration equipment.<br />

Ultrasonic wave speeds were also measured <strong>and</strong> the static bending <strong>MOE</strong> was determined by<br />

using a four point bending machine. The samples were then dried to 12% moisture content<br />

<strong>and</strong> the described tests were repeated <strong>for</strong> dry specimens <strong>and</strong> at the end the failure strength was<br />

determined by four point bending equipment. The result of this research showed that the<br />

relationship between dry static bending <strong>MOE</strong> versus green stress wave velocity or the<br />

corresponding green <strong>MOE</strong> can directly be used to predict the dry static bending <strong>MOE</strong> (Ibid).<br />

In a study per<strong>for</strong>med in New Zeal<strong>and</strong> in 1998, 300 pine logs (species not reported but likely<br />

to be radiata pine) were tested using acoustics (Tsehaye et al, 2000). One end of each log was<br />

hit with a hammer containing an accelerometer that records the moment of impact. The sound<br />

wave passed along the 4.2 m long log <strong>and</strong> its arrival at the other end detected by a second<br />

accelerometer that was pressed against the log end. The 27-year-old pine logs from Mamaku<br />

Plateau in the Central North Isl<strong>and</strong> were sorted into three groups (27, 39, 27 logs), according<br />

to the speed of sound. After the measurements, the logs were sawn into 100 x 40 mm boards<br />

<strong>and</strong> then kiln dried to 12% moisture content. The <strong>MOE</strong>s of the dressed boards were<br />

determined by a stress-grading machine. The regression between the squared velocity of<br />

sound <strong>and</strong> the mean modulus of elasticity (300 logs) gave an R 2 of 0.46. Within the single<br />

groups the R 2 was between 0.45 <strong>and</strong> 0.57.<br />

From these results it follows, that acoustic sorting of logs provides the opportunity to send<br />

only the best quality, high stiffness logs, to the sawmill (Tsehaye et al, 2000). In this study<br />

only two parameters, speed of sound <strong>and</strong> <strong>MOE</strong> provided by a stress-grading machine were<br />

compared. The effective strength <strong>and</strong> stiffness of the boards was not determined.<br />

12


Ross et al (2005) conducted an investigation evaluating Douglas fir (Pseudotsuga menziesii)<br />

peeler cores. The longitudinal stress wave method was used to evaluate peeler cores from 111<br />

Douglas fir stems. Then the 2.6 m long peeler cores were sawn into 90 x 40 mm boards. The<br />

dynamic <strong>MOE</strong> of each board was determined in green, as well as dry condition (6-7% MC)<br />

by using stress wave <strong>and</strong> transversal vibration techniques. Afterwards, the boards were tested<br />

to failure strength in bending to determine the static bending <strong>MOE</strong> <strong>and</strong> the modulus of<br />

rupture. This study showed that stress-wave-predicted <strong>MOE</strong> of peeler cores is a good<br />

predictor of both dynamic <strong>and</strong> static <strong>MOE</strong> of timber obtained from the cores. Very strong<br />

relationships were found between stress wave <strong>and</strong> <strong>MOE</strong> of the peeler cores <strong>and</strong> dynamic<br />

<strong>MOE</strong> (stress wave <strong>MOE</strong> <strong>and</strong> vibration <strong>MOE</strong>) as well as static <strong>MOE</strong> (bending <strong>MOE</strong> <strong>and</strong><br />

tensile <strong>MOE</strong>) of the timber. However, the correlations between stress wave <strong>MOE</strong> of the<br />

peeler cores <strong>and</strong> the bending <strong>and</strong> tensile strength (<strong>MOR</strong>) were low. This investigation has<br />

shown that a basic grading of specimens is possible using stress wave techniques (Ibid).<br />

Grabianowski et al (2006) per<strong>for</strong>med a study about acoustic measurements on st<strong>and</strong>ing trees,<br />

logs <strong>and</strong> green timber of young Pinus radiata trees (aged 8 to 11-years-old). Thereby a time<br />

of flight tool (Fakopp ® 2D) <strong>and</strong> a resonance based system (WoodSpec) were used to evaluate<br />

the stiffness of the test material. The aim of this study was to determine how well acoustic<br />

measurements of green logs, estimate the properties of timber cut from those logs. The<br />

correlation using Fakopp ® 2D between the log values <strong>and</strong> those <strong>for</strong> the average of two boards<br />

from each log was 0.94. By using the resonance based system WoodSpec this relationship<br />

was 0.86. Significant correlations were found between the two tools, especially <strong>for</strong> stems (R 2<br />

= 0.96) (Grabianoski et al, 2005).<br />

During Combigrade Phase 1, Hanhijärvi et al reviewed the results from five previous<br />

investigations into NDT <strong>for</strong> strength prediction published between 1984 <strong>and</strong> 1997. Based on<br />

this review it was concluded that:<br />

• Correlations (coefficient of determination, R 2 ) varied between studies, probably due to<br />

differing materials <strong>and</strong> methods.<br />

• The highest correlation by any parameter tested achieved R 2 =0.7.<br />

• <strong>MOE</strong> is the best single variable <strong>for</strong> prediction of strength, followed by KAR <strong>and</strong><br />

density.<br />

• A combination of predictors provides greater accuracy than a single predictor.<br />

Other relevant findings in the literature included:<br />

• Görlacher (1984) found that the natural frequency (dynamic <strong>MOE</strong>) correlated well<br />

with static test results, as did Blass <strong>and</strong> Gard (1994) in their tests on Douglas fir.<br />

• In separate studies S<strong>and</strong>oz (1989) <strong>and</strong> Diebold et al (2000) found R 2 =0.45 <strong>and</strong> 0.53<br />

respectively <strong>for</strong> ultrasonic speed <strong>and</strong> strength.<br />

• On a small number of specimens Oja et al (2000) found a prediction of R 2 =0.41 <strong>for</strong> Xray<br />

(density <strong>and</strong> knot volume) <strong>and</strong> strength of sawn boards.<br />

• Similarly to the Combigrade project, Fonselius et al (1997) found that the accuracy of<br />

the predictors varies between species. For example in the Fonselius study it was found<br />

that knots explained 57% of the strength of pine compared to 27% <strong>for</strong> spruce.<br />

For Phase 1 of the Combigrade project (Hanhijärvi et al, 2005) approximately 100 logs each<br />

of spruce (Picea spp.) <strong>and</strong> pine (Pinus spp.) were investigated by testing them with different<br />

non-destructive testing methods. The logs were scanned by X-ray equipment <strong>and</strong> natural<br />

frequency <strong>and</strong> acoustic tomography measurements (only on 75 spruce logs, no pine logs),<br />

be<strong>for</strong>e they were sawn into nominal 150 x 50 mm boards. Approximately one board per log<br />

13


was selected <strong>and</strong> the batch was dried to an average of 10%. Final dimensions after drying <strong>and</strong><br />

dressing were 146 x 46 mm, with a pencil round profile.<br />

A range of data was collected <strong>for</strong> the trial boards:<br />

• scanned <strong>for</strong> digital image analysis;<br />

• natural frequency measurements twice (two different operators);<br />

• X-ray scans;<br />

• acoustic-ultrasonic measurements;<br />

• sloping grain;<br />

• density by gross measure <strong>and</strong> by gamma ray;<br />

• Finnograder (gamma rays, microwaves, infrared radiation to predict strength);<br />

• Raute Timgrader <strong>MOE</strong>;<br />

• Compression <strong>MOE</strong>;<br />

• Average annual ring width by image analysis.<br />

The test material was loaded to failure in bending, <strong>and</strong> modulus of elasticity, bending<br />

strength, knot area <strong>and</strong> density were measured to determine grade.<br />

Findings from Phase 1 of the Combigrade project are summarised below:<br />

• Spruce <strong>and</strong> pine populations behave differently in regard to the predictors;<br />

• Stiffness parameters had the best single-variable predictions of bending strength:<br />

<strong>MOE</strong> measured by either static method, vibration method or by ultrasonic method.<br />

• Correlations between NDT density measurements (gross measurements, X-ray or<br />

gamma ray) <strong>and</strong> strength were within a similar range with X-ray providing the best<br />

value.<br />

• Density is a better grade predictor <strong>for</strong> pine than <strong>for</strong> spruce.<br />

• Knot parameters provide good predictions of strength <strong>and</strong> density <strong>for</strong> pine, but not<br />

spruce<br />

• Irradiation equipment (X-ray <strong>and</strong> gamma ray) provide slightly better strength<br />

prediction than surface inspection such as KAR.<br />

• Sloping grain measurements didn’t have the potential to predict strength.<br />

• Relatively strong correlations were obtained from log measurements with destructive<br />

board tests.<br />

• Log X-ray <strong>and</strong> dynamic <strong>MOE</strong> based on natural frequency both provide R 2 of 0.60 or<br />

better <strong>for</strong> pine.<br />

• Combinations of devices provided correlations of R 2 =0.80 to 0.85 <strong>for</strong> pine <strong>and</strong> 0.60 to<br />

0.65 <strong>for</strong> spruce.<br />

• Combination of knot measurements with density <strong>and</strong> annual ring width provides<br />

effective predictions <strong>for</strong> strength with R 2 =0.7 (pine) <strong>and</strong> 0.6 (spruce).<br />

Based on recommendations from Phase 1 of the Combigrade project, a larger sample with<br />

replicates of different sectional sizes was tested to <strong>for</strong>m the Phase 2 follow up trial<br />

(Hanhijärvi et al, 2008). For Phase 2 more than 1000 logs each of spruce <strong>and</strong> pine were<br />

sampled <strong>and</strong> after NDT data gathering from the logs, one board per log was selected from a<br />

mix of seven different size classes. In addition to the larger sample <strong>and</strong> range of dimensions,<br />

Phase 2 boards weren’t planed after kiln drying. Results determined during Phase 2 of the<br />

Combigrade project can be summarized as:<br />

• The two species behaved differently, similar to the small sample tested during Phase<br />

1, with stronger correlation between predictors <strong>and</strong> strength usually achieved by pine.<br />

14


• Cross-section dimensions also behaved differently, providing an insight to how size<br />

affects correlations.<br />

• The correlation of density to <strong>MOE</strong> <strong>and</strong> <strong>MOR</strong> in bending decreases with an increase in<br />

section size.<br />

• The results <strong>and</strong> conclusions <strong>for</strong> correlations’ analyses were similar to the results from<br />

the smaller sample tested in Phase 1.<br />

St<strong>and</strong>ard requirements<br />

The Australian St<strong>and</strong>ard AS1720.1:1997 Timber structures, Part 1: Timber properties defines<br />

structural per<strong>for</strong>mance of machine <strong>graded</strong> pine (MGP) by the 5th percentile strength (<strong>MOR</strong>)<br />

of a sample population (r<strong>and</strong>om position tested) <strong>and</strong> by an average <strong>MOE</strong> (St<strong>and</strong>ards<br />

Australia, 1997). In the current AS/NZS4063, the design properties are related to<br />

characteristic strength <strong>and</strong> <strong>MOE</strong>:<br />

• characteristic strength is lower 75% Confidence limit of the 5 th percentile strength of<br />

the population estimated from tests on a sample;<br />

• characteristic <strong>MOE</strong> is 75% Confidence limit of the average population <strong>MOE</strong><br />

estimated from tests on a sample.<br />

To achieve these requirements, MGP material needs to be <strong>graded</strong> on the basis of<br />

AS/NZS1748:2003 Timber-Stress-<strong>graded</strong>-Product requirements <strong>for</strong> mechanically stress<strong>graded</strong><br />

timber. This St<strong>and</strong>ard requires that every board is initially <strong>graded</strong> by a mechanical<br />

stress-grader which sorts the timber on the basis of its <strong>MOE</strong> (low point or mean or<br />

combination) <strong>and</strong> secondly by a visual inspection to detect strength-limiting <strong>and</strong> utilitylimiting<br />

characteristics.<br />

The machine stress-grader settings are determined by continuous monitoring (based on<br />

r<strong>and</strong>om position testing reference) of the values obtained from batches. Strength-limiting<br />

parameters are knots, resin- <strong>and</strong> bark pockets, cross- <strong>and</strong> heart shakes <strong>and</strong> splits. Utilitylimiting<br />

parameters <strong>for</strong> softwood species include dimensional tolerance, squareness, knots,<br />

wane <strong>and</strong> want, machine skip, <strong>and</strong> distortion (bow/spring/twist). In the visual over-ride, high<br />

attention must be paid to the board ends, where the machine stress grader cannot determine<br />

the <strong>MOE</strong> (St<strong>and</strong>ards Australia, 2003).<br />

The European st<strong>and</strong>ard EN14081-1:2003 Timber structures – Strength <strong>graded</strong> structural<br />

timber with rectangular cross section – Part 1: General requirements requires either visual or<br />

machine <strong>graded</strong> timber. Visually <strong>graded</strong> timber accounts <strong>for</strong>:<br />

• strength-reducing characteristics like knots, slope of grain, density, rate of growth <strong>and</strong><br />

fissures;<br />

• geometric characteristics like wane <strong>and</strong> warp; <strong>and</strong><br />

• biological characteristics like fungal <strong>and</strong> insect damage.<br />

If the timber is machine <strong>graded</strong> the boards have to be <strong>graded</strong> on their full length. If that is not<br />

the case, as in bending type machines, the non-<strong>graded</strong> part needs to be visually examined (EN<br />

14081-1, 2003) as <strong>for</strong> the Australian St<strong>and</strong>ard.<br />

The European st<strong>and</strong>ard requires a biased position test as the reference static bending test.<br />

15


Study Methodology<br />

Materials<br />

Sampling method<br />

Logs provided <strong>for</strong> the trials were selected from a mix of butt logs (30%) <strong>and</strong> upper logs (70%)<br />

in order to over-sample the low grades <strong>and</strong> ensure sufficient representation of the grades<br />

below MGP10 <strong>and</strong> in the lower half of the MGP10 population. Rather than sample 1<br />

board/log <strong>and</strong> have a large number of individual logs, this project aimed to maximize the<br />

number of boards recovered from each log to provide the best experiment <strong>for</strong> comparing<br />

results from the NDT log tests <strong>and</strong> the subsequent results from board testing.<br />

Paper log end templates based on Smith et al (2003) were adhered to both ends of each log to<br />

enable identification from log (Figure 1) through the green <strong>and</strong> dry chains <strong>and</strong> subsequent<br />

testing. In<strong>for</strong>mation provided on each label enabled the unique log number to remain on sawn<br />

boards through all stages of processing <strong>and</strong> testing as well as determining the log end (small<br />

end or large end) <strong>and</strong> relative in-log position of each board.<br />

Figure 1. Log end template.<br />

Stage 1 Stiffness limited radiata pine (Radiata E)<br />

Stiffness limited Pinus radiata, radiata pine was sourced from 67 logs from two plantations<br />

located near the Victoria-New South Wales border <strong>and</strong> aged approximately 28-years-old. The<br />

average centre diameter <strong>for</strong> the batch was 370 mm (range 280 mm to 500 mm). From this<br />

batch the target number of board samples was 600. The sawmilling process provided 992<br />

green boards, of which complete data sets <strong>for</strong> all trials were compiled <strong>for</strong> 517 boards. An<br />

overview of the testing conducted on this batch is provided in Figure 2.<br />

16


stage 1<br />

Carter Holt Harvey Myrtle<strong>for</strong>d<br />

DPI&F<br />

Brisbane<br />

DPI&F<br />

Brisbane<br />

67 radiata pine logs<br />

Three vibration measurements<br />

Assortment of the boards<br />

52 x 100 mm 266 pieces<br />

42 x 100 mm 548 pieces<br />

42 x 80 mm 9 pieces<br />

Sawing the logs <strong>and</strong> numbering the boards<br />

25 x 150 mm 7 pieces<br />

25 x 100 mm 153 pieces<br />

25 x 80 mm 9 piecess<br />

Bing® vibration measurements on green boards<br />

Determination of <strong>MOE</strong> <strong>and</strong> a range of signal descriptors<br />

Docked all the boards to a maximum length of 5.0 meters<br />

<strong>and</strong> resaw the 52 mm boards to 42 mm to simplify further<br />

tests<br />

R<strong>and</strong>om sub-sample of 600 boards (42 x 100 mm)<br />

Weyerhaeuser Caboolture<br />

Hyne Timber<br />

Melawondi<br />

DPI&F Brisbane<br />

17<br />

Dimensions <strong>and</strong> density measurements by gamma ray<br />

<strong>MOE</strong> by Thumper in-line vibration system<br />

Drying <strong>and</strong> dressing the boards (? 12% moisture content / 35 mm by 90 mm)<br />

Metriguard (MSR) testing at Weyerhaeuser<br />

Determination of <strong>MOE</strong> along the boards<br />

WoodEye scanner<br />

Determination of external defects<br />

Bing® vibration measurements on dry board<br />

Determination of <strong>MOE</strong> <strong>and</strong> a range of signal descriptors<br />

Four point reference bending test at Salisbury (Biased & R<strong>and</strong>om)<br />

Determination of <strong>MOE</strong> & <strong>MOR</strong><br />

Figure 2. Sample flow <strong>for</strong> Stage 1, stiffness-limited radiata pine.<br />

Stage 2 Caribbean pine (Caribbean)<br />

One-hundred <strong>and</strong> sixteen logs representing the sub-tropical exotic pine resource of south-east<br />

Queensl<strong>and</strong> were provided <strong>for</strong> the trial. The age of the source plantation was not provided by<br />

the supplier who estimated that the trees were 25- to 30-years old. The average log diameter<br />

<strong>for</strong> this batch was 250 mm (range 230 mm to 270 mm). The target number of boards from this<br />

sample was 400 <strong>and</strong> from 741 green boards, 489 were used to provide the data <strong>for</strong> the batch.<br />

An overview of the testing is provided in Figure 3 below.<br />

Stage 2<br />

QPIF Brisbane, Weyerhaeuser Caboolture<br />

QPIF<br />

Brisbane<br />

106 Caribbean pine logs<br />

Three vibration measurements Drying <strong>and</strong> dressing the boards (? 12% moisture content / 35 mm by 90 mm)<br />

Assortment of the boards<br />

90 x 19 mm 106 pieces<br />

75 x 42 mm 40 pieces<br />

100 x 40 mm 595 pieces<br />

Sawing the logs <strong>and</strong> numbering the boards<br />

Bing® vibration measurements on green boards<br />

Determination of <strong>MOE</strong> <strong>and</strong> a range of signal descriptors<br />

Figure 3. Sample flow <strong>for</strong> Stage 2, Caribbean pine.<br />

Weyerhaeuser Caboolture<br />

QPIF<br />

Brisbane<br />

Hyne<br />

Timber<br />

Melawondi<br />

QPIF<br />

Brisbane<br />

Metriguard (MSR) testing at Weyerhaeuser<br />

Determination of <strong>MOE</strong> along the boards<br />

Bing® vibration measurements on dry board<br />

Determination of <strong>MOE</strong> <strong>and</strong> a range of signal descriptors<br />

WoodEye® scanner<br />

Determination of external defects<br />

Four point reference bending test at Salisbury (Biased & R<strong>and</strong>om)<br />

Determination of <strong>MOE</strong> & <strong>MOR</strong><br />

Stage 3 Strength limited radiata pine (Radiata R)<br />

Forty-seven logs with an average centre diameter of 420 mm (range 300 mm to 580 mm)<br />

representing the strength-limited radiata pine resource of Western Australia were provided <strong>for</strong><br />

testing with a target sample of 400 dried boards. After sawing <strong>and</strong> drying, 582 boards<br />

provided the data set <strong>for</strong> this resource sample. The trial flow is depicted in Figure 4.


Stage 3<br />

Wespine Bunbury<br />

Wespine<br />

Bunbury<br />

47 radiata pine logs<br />

Three vibration measurements Drying <strong>and</strong> dressing the boards (? 12% moisture content / 35 mm by 90 mm)<br />

Assortment of the boards<br />

100 x 40 mm 642 boards<br />

Sawing the logs <strong>and</strong> numbering the boards<br />

Bing® vibration measurements on green boards<br />

Determination of <strong>MOE</strong> <strong>and</strong> a range of signal descriptors<br />

Wespine Bunbury<br />

DPI&F<br />

Brisbane<br />

Hyne<br />

Timber<br />

Melawondi<br />

DPI&F<br />

Brisbane<br />

18<br />

Metriguard (MSR) testing at Wespine<br />

Grain angle, moisture content, <strong>and</strong> unsound wood position measurements by X-Ray<br />

Determination of <strong>MOE</strong> along the boards<br />

Bing® vibration measurements on dry board<br />

Determination of <strong>MOE</strong> <strong>and</strong> a range of signal descriptors<br />

WoodEye scanner<br />

Determination of external defects<br />

Figure 4. Overview of testing, Stage 3, strength-limited radiata pine.<br />

Four point reference bending test at Salisbury (Biased & R<strong>and</strong>om)<br />

Determination of <strong>MOE</strong> & <strong>MOR</strong><br />

Kiln drying<br />

After completion of tests in the green (unseasoned) condition, boards were transported to a<br />

commercial softwood plant <strong>for</strong> high temperature kiln drying to a target average moisture<br />

content of 12%. After equalisation, the test boards were dressed to 90 x 35 mm.<br />

Equipment <strong>and</strong> methods<br />

Non destructive testing- mechanical properties measurement<br />

Machine stress rating<br />

Metriguard Continuous Lumber Tester (CLT 7100) equipment was used to collate <strong>MOE</strong><br />

profiles <strong>for</strong> Stage 1 (radiata E) <strong>and</strong> Stage 2 (Caribbean) dry boards <strong>and</strong> Metriguard High<br />

Speed Continuous Lumber Tester (HCLT) <strong>for</strong> the Stage 3 (radiata R) dry boards in order to<br />

prescribe stress ratings. The boards were bent flatwise by rollers downward <strong>and</strong> then upward<br />

as depicted in Figure 5 below.<br />

Figure 5. Schematic diagram of Metriguard CLT (source: Web 4)<br />

Thereby the bending <strong>for</strong>ce <strong>and</strong> the deflection in both bending sections were measured. The<br />

local <strong>MOE</strong>s at intervals of 13.9 mm were automatically calculated, from which the average<br />

<strong>and</strong> the low point <strong>MOE</strong> were provided on the full length (excluding the leading <strong>and</strong> trailing<br />

820 mm ends sections of boards). Additionally, the Metriguard <strong>MOE</strong> profile <strong>for</strong> the static test<br />

section was extracted to allow correlation with the static bending <strong>MOE</strong> <strong>and</strong> <strong>MOR</strong> test results.


The predictors used <strong>for</strong> the analysis were:<br />

• full board profile extraction: Metri_avg board <strong>and</strong> Metri_min board;<br />

• static bending span: Metri_avg Static Test <strong>and</strong> Metri_min Static Test.<br />

In-line acoustic test<br />

The Thumper is an in-line measurement device used to capture the longitudinal frequency <strong>and</strong><br />

combine the data with parameters captured by the gamma-ray <strong>and</strong> laser measuring tools in the<br />

same green chain to calculate the dynamic <strong>MOE</strong> of each green board. It consists of a<br />

mechanical pneumatic cylinder that strikes the board end <strong>and</strong> a microphone which receives<br />

the signal. In front of the cylinder is a laser proximity switch which senses the presence of the<br />

board <strong>and</strong> triggers the firing of the cylinder to strike the board. The impact sets up a stress<br />

wave which travels immediately through the board <strong>and</strong> is detected by the microphone. With<br />

the vibration fundamental frequency collected by the microphone <strong>and</strong> the dimensions as well<br />

as the density provided by the gamma ray system, the dynamic <strong>MOE</strong> of each board can be<br />

calculated, allowing sorting of the green boards into stiffness classes.<br />

The in-line Thumper system delivers a vibration <strong>MOE</strong> parameter (Thumper_<strong>MOE</strong>).<br />

Off-line acoustic test<br />

The acoustic velocity <strong>and</strong> vibration measurements were captured using Bing ® <strong>and</strong> WISIS<br />

(Wood in Situ Inspection) products which were developed by CIRAD (http://www.xylometry.org/en/softwares.html).<br />

WISIS measures the time of flight from an induced stress wave, based on a time-frequency<br />

analysis, not on a frequency analysis like the Bing ® procedure. From the extracted results the<br />

<strong>MOE</strong> can be calculated by using Equation 1. This type of measurement can also be applied to<br />

st<strong>and</strong>ing trees to predict wood stiffness similar to commercial tools currently available in the<br />

<strong>for</strong>est wood quality sector. With velocity measurement, no boundary conditions are required,<br />

contrary to vibration analysis.<br />

Bing ® allows determination of the bending <strong>and</strong> compression <strong>MOE</strong> by analysis of the natural<br />

vibration spectrum of a piece of wood. The technique is also known as the resonance method<br />

as it allows determination of resonance frequencies of a beam from its response to an impact.<br />

The two systems consist of the devices shown below in Figure 6 <strong>and</strong> associated software.<br />

19


F<br />

G<br />

C<br />

D<br />

Figure 6. Equipment used <strong>for</strong> the acquisition of the off-line acoustic signals.<br />

A Laptop or Desktop<br />

B Data acquisition card (Pico Technology Type Picoscope 3224)<br />

C Accelerometer (Brüel & Kjaer Type 4397)<br />

D Impact hammer (Brüel & Kjaer Type 8206-2)<br />

E Conditioner (Endevco Type 4416B) <strong>for</strong> accelerometer <strong>and</strong> impact hammer<br />

F Low pass filter to avoid aliasing effect<br />

G Screw <strong>for</strong> the accelerometer attachment (magnet system)<br />

Logs <strong>and</strong> green boards were tested using the Bing ® <strong>for</strong>:<br />

• longitudinal vibration-<br />

o compression waves,<br />

o frequency 2000 Hz,<br />

o four resonance frequencies viz F1 to F4, <strong>and</strong><br />

• transverse vibration-<br />

o flexion waves,<br />

o frequency 1000 Hz,<br />

o five resonance frequencies viz F1 to F5.<br />

A<br />

B<br />

These measurements provided the following range of vibration signal descriptors: the <strong>MOE</strong>s<br />

in each configuration were calculated according to the equations described in the paragraph<br />

“Resonance method” above.<br />

The vibration spectra were analysed in order to provide the following range of vibration<br />

signal descriptors:<br />

• dynamic <strong>MOE</strong> associated with Fn (<strong>MOE</strong>_n).<br />

• spectral centre of gravity divided by the fundamental frequency, F1 in %<br />

(SCGravity).<br />

• spectral extent of gravity divided by the fundamental frequency, F1 in %<br />

(SB<strong>and</strong>width).<br />

• spectral slope divided the fundamental frequency, F1 in % (SSlope).<br />

• quality factor (inverse internal damping or friction) associated with Fn (FacQn).<br />

20<br />

E<br />

E


• inharmonicity factor (IF).<br />

• mean power in frequency sub-b<strong>and</strong>; between 0 <strong>and</strong> f1; f1 <strong>and</strong> f2... (MSPower).<br />

• sub-b<strong>and</strong> energy ratio, between the resonance frequencies, i.e. sub-b<strong>and</strong> energy/total<br />

range energy (MSNrjRatioF_n).<br />

• mean power of the sub-b<strong>and</strong>, resonance shoulder defined by a pass b<strong>and</strong> of 20dB<br />

(Pow_n).<br />

• sub-b<strong>and</strong> energy ratio, resonance shoulder defined by a pass b<strong>and</strong> of -20dB<br />

(MSNrjRatioF_n).<br />

• <strong>MOE</strong> extraction using Timoshenko’s model <strong>and</strong> Bordonné’s solution (<strong>MOE</strong>T).<br />

A full description of the parameters is provided in Appendix 1.<br />

Maximal relative error on vibration <strong>MOE</strong>s<br />

After the signal conversion from analogue to digital, the Eigen frequencies were calculated by<br />

means of the Fast Fourier Trans<strong>for</strong>m. The resolution (r) is a function of sampling frequency<br />

(Fe) <strong>and</strong> of the number of measurement points (N):<br />

f e r = (Equation 7)<br />

N<br />

The absolute error Δf can be increased by the experimental absolute error (1 Hz) <strong>and</strong> half of<br />

the resolution. There<strong>for</strong>e the maximal relative error <strong>for</strong> Bing ® measurements can be<br />

determined as follows:<br />

Compression:<br />

Flexion:<br />

Δf Δf<br />

i i _ exp + 0.<br />

5 ⋅r<br />

=<br />

f f<br />

i<br />

Resolution r = 1.19 Hz<br />

Δf i _ exp = 1.<br />

0 Hz<br />

i<br />

(Equation 8)<br />

Δf<br />

1<br />

f1 = 193 Hz = 0.<br />

83%<br />

f<br />

Δf<br />

2<br />

f2 = 583 Hz = 0.<br />

27 %<br />

f<br />

Δf<br />

3<br />

f3 = 869 Hz = 0.<br />

18 %<br />

f<br />

Δf<br />

4<br />

f4 = 1161 Hz = 0.<br />

14 %<br />

f<br />

Resolution r = 0.3 Hz<br />

Δf i _ exp = 1.<br />

0 Hz<br />

1<br />

2<br />

3<br />

4<br />

21


Δf<br />

1<br />

f1 = 10.4 Hz = 11.<br />

1%<br />

f<br />

Δf<br />

2<br />

f2 = 22.6 Hz = 5.<br />

1%<br />

f<br />

Δf<br />

3<br />

f3 = 54.2 Hz = 2.<br />

1%<br />

f<br />

Δf<br />

4<br />

f4 = 92.1 Hz = 1.<br />

2 %<br />

f<br />

1<br />

2<br />

3<br />

4<br />

The maximal relative error of the dynamic <strong>MOE</strong> can be calculated by the following equation:<br />

Δ<strong>MOE</strong><br />

<strong>MOE</strong><br />

Bing , L<br />

Bing , L<br />

Δm<br />

Δh<br />

Δb<br />

ΔL<br />

Δf<br />

1<br />

= + + + + 2<br />

m h b L f<br />

22<br />

1<br />

(Equation 9)<br />

For the mass: m = 5.4 ± 0.02 kg For the width: b = 30 ± 1 mm<br />

For the height: h = 90 ± 1 mm For the length: L = 5000 ± 20 mm<br />

Thus the maximal relative errors on compression measurements are:<br />

Δ<strong>MOE</strong><br />

<strong>MOE</strong><br />

Δ<strong>MOE</strong><br />

<strong>MOE</strong><br />

Bing , L , 1<br />

Bing , L , 1<br />

Bing , L , 3<br />

Bing , L , 3<br />

=<br />

=<br />

6.<br />

4 %<br />

5.<br />

1%<br />

Δ<strong>MOE</strong>Bing<br />

, L , 2<br />

, = 5.<br />

3%<br />

,<br />

<strong>MOE</strong><br />

Bing , L , 2<br />

Δ<strong>MOE</strong>Bing<br />

, L , 4<br />

, = 5%<br />

<strong>MOE</strong><br />

Bing , L , 4<br />

For <strong>MOE</strong>Bing, F, 1-4 the maximal relative errors are:<br />

Δ<strong>MOE</strong><br />

<strong>MOE</strong><br />

Δ<strong>MOE</strong><br />

<strong>MOE</strong><br />

Bing , F , 1<br />

Bing , F , 1<br />

Bing , F , 3<br />

Bing , F , 3<br />

=<br />

27%<br />

Δ<strong>MOE</strong>Bing<br />

, F , 2<br />

, = 15%<br />

,<br />

<strong>MOE</strong><br />

Bing , F , 2<br />

Δ<strong>MOE</strong>Bing<br />

, F , 4<br />

= 9%<br />

, = 7%<br />

<strong>MOE</strong><br />

Bing , F , 4<br />

The data processing of the acquired digital signal involved the use of zero padding <strong>and</strong><br />

smoothing procedures, to provide a more accurate reading of the resonance frequency.<br />

Gamma ray<br />

An in-line gamma-ray system (Geological & Nuclear sciences Ltd) was used to capture<br />

estimated density <strong>and</strong> moisture content data. The boards move transversely <strong>and</strong> are pushed<br />

along by lugs at 450 mm intervals on a chain conveyer system. The speed of the system can<br />

be varied up to 100 boards/minute, but is typically around 70-80 boards/minute. A system of<br />

four laser range instruments, positioned above <strong>and</strong> below the line, is used to measure the<br />

thickness, width <strong>and</strong> the length of the piece of timber. This in<strong>for</strong>mation is combined with the<br />

gamma ray result to provide the average green density of the board. Gamma ray<br />

measurements are made simultaneously at four positions along the board by using four<br />

separate heads positioned between the chains. These measurements are combined to give


average values which are used to sort the boards <strong>for</strong> subsequent kiln drying or machine stress<br />

grading.<br />

Non destructive testing- structural properties measurement<br />

Linear high grader (LHG)<br />

The LHG was developed by Coe Newnes/McGehee ULC (Canada). Coe Newnes/McGehee<br />

ULC provides solutions <strong>for</strong> sawmill systems, planer mill systems, scanning <strong>and</strong> optimization<br />

technology (Web 5).<br />

The functions of the LHG are summarised as:<br />

• Measures wane, skip <strong>and</strong> holes.<br />

• Measures grain angle, moisture content, <strong>and</strong> unsound wood locations.<br />

• Determines best grade/length based on mill’s grading rules.<br />

• Marks board after scanner – ID number.<br />

• Read boards ID.<br />

• Sends trim <strong>and</strong> grade decision.<br />

The LHG utilises X-ray technology to analyse density variation allowing the identification of<br />

knots within boards. Data are combined with low-point <strong>MOE</strong> results determined by<br />

Metriguard HCLT equipment to predict <strong>MOR</strong>. The LHG scanning was only used on the<br />

Radiata R resource.<br />

LHG KAR is the LHG’s estimation of knot area ratio (KAR). A second parameter was<br />

extracted from the knot position data obtained with the X-ray, called LHG I ratio which is<br />

the ratio of the inertia (I value) of the knots in the window to the inertia of the full cross<br />

section. The maximum value of I ratio was extracted using a Gaussian sliding-window.<br />

LHG Weighted KAR is the same as LHG KAR, but only <strong>for</strong> knots in the outer quartile of<br />

the depth of the beam, where the outer quartile has weighting of 1.0 <strong>and</strong> the inner quartiles<br />

have weighting of 0.1, again using a Gaussian sliding-window.<br />

LHG prediction of strength (LHG predicted Strength) was derived from the combination of<br />

LHG parameters <strong>and</strong> Metriguard HCLT profile from the biased <strong>and</strong> r<strong>and</strong>om test span. All<br />

these parameters were extracted locally from the data gathered from the static test span.<br />

WoodEye ®<br />

A WoodEye ® optical scanning system was used to record defect type, size <strong>and</strong> location<br />

features. The WoodEye ® scanning system used combines:<br />

• grey scale camera;<br />

• RGB colour camera;<br />

• tracheid effect LASER; <strong>and</strong><br />

• profile detection LASER.<br />

The sensor-system scans each piece on all four sides <strong>for</strong> natural features such as black knot,<br />

sound knot, fibre knot as well as board geometry. Other features such as wane <strong>and</strong> pith<br />

weren’t considered during this study.<br />

For strength grading the knot categories are the most interesting. Black knot (Bk) <strong>and</strong> Sound<br />

knot (Sk) refers to knot defects created from the grey scale sensor. Fibre knot (Fk) is created<br />

from the laser sensor <strong>and</strong> is based on detection of a fibre disturbance. There<strong>for</strong>e, a certain<br />

23


knot can give rise to both a Fibre knot <strong>and</strong> a Black/Sound knot. The sizes can differ due to the<br />

detection situation. Fibre knots tend to be larger. The contour of the combined knot silhouette<br />

(WE) was used as a predictor.<br />

The beam profile construction is based on a rendering discrete process over the length of the<br />

beam with 1 cm steps. Each profile step is associated with a defects’ projection of a portion on<br />

the length of the beam equivalent to two times the height of the beam. The projection window<br />

can be:<br />

• rectangular on the entire beam portion;<br />

• rectangular external: 1/3 of the top <strong>and</strong> bottom part of the beam height;<br />

• rectangular interior: 1/3 of the centre part of the beam height;<br />

• triangular according to a diamond-shaped pattern.<br />

All the profiles are filtered with a Blackman sliding-window of which the size is equivalent to<br />

the beam height. Descriptions of the WoodEye ® profiles are provided in Appendix 2.<br />

Destructive st<strong>and</strong>ard static bending tests<br />

Static bending tests were per<strong>for</strong>med using a testing method in accordance with AS/NZS<br />

4063:1992. Biased position tests were used on every board, <strong>and</strong> where the r<strong>and</strong>om position<br />

test location was within the useful remnant of the board after biased testing, a r<strong>and</strong>om position<br />

test was also per<strong>for</strong>med on the same board. <strong>MOE</strong> <strong>and</strong> <strong>MOR</strong> were calculated <strong>for</strong> each test.<br />

Protocols used <strong>for</strong> the selection of biased <strong>and</strong> r<strong>and</strong>om samples are described in Appendix 3.<br />

The load <strong>for</strong> the reference testing was applied <strong>and</strong> measured with a Shimadzu UDH-30 tonne<br />

(300 kN) universal testing machine depicted below (Figure 7).<br />

Figure 7. Shimadzu UDH-30 t Universal testing machine.<br />

The term ‘universal’ indicates that the machine is capable of per<strong>for</strong>ming a range of tests<br />

including static bending, tension, compression, shear <strong>and</strong> hardness tests on large samples. The<br />

support consists of a solid steel roller 240 mm long by 50 mm diameter <strong>and</strong> a flat mounting<br />

plate. The mounting plates have two holes which are used to locate the plate over machine<br />

24


olts situated in the centre of the supports. The roller, <strong>and</strong> to a lesser extent the mounting<br />

plate, are held in place by two springs which attach to bolts protruding from the centre of the<br />

load roller. The Shimadzu UDH-30 is a ‘Grade A’ testing machine in accordance with<br />

Australian St<strong>and</strong>ard AS 2193:2002 Calibration <strong>and</strong> classification of <strong>for</strong>ce-measuring systems.<br />

The Australian <strong>and</strong> New Zeal<strong>and</strong> St<strong>and</strong>ard, AS/NZS 4063:1992 Timber-Stress-<strong>graded</strong>-Ingrade<br />

strength <strong>and</strong> stiffness evaluation, requires that the specimens shall be conditioned to a<br />

temperature of 20 ±3°C <strong>and</strong> in an environment having a relative humidity of 65 ±5%. This<br />

conditioning shall continue until the moisture content is stable within each piece (10-15%).<br />

Moisture content was checked using an electric resistance moisture meter to confirm that the<br />

boards had conditioned to the range specified in the st<strong>and</strong>ard. Bending strength <strong>and</strong> stiffness<br />

were then tested according to the methods specified in AS/NZS 4063:1992.<br />

In the middle of the span the deflection was measured with a strain gauge type linear<br />

displacement transducer. The bending test span was 1620 mm with load applied at four points<br />

<strong>and</strong> the span-to-depth ratio was 18:1. The load deflection curve was measured up to 1.6 kN<br />

<strong>for</strong> all specimens. The modulus of elasticity can be determined from the slope of the linear<br />

relationship between the applied load (P) <strong>and</strong> the resulting deflection (ε) using the following<br />

equation:<br />

3<br />

23 l s ΔP<br />

<strong>MOE</strong> = ⋅ ⋅ 3<br />

108 bh ε<br />

25<br />

(Equation 10)<br />

After removal of the displacement transducer, loading continued until failure of the specimen.<br />

The modulus of rupture was calculated using the following equation:<br />

l P<br />

bh<br />

s <strong>MOR</strong> = 2<br />

(Equation 11)<br />

The following equation gives an estimation of the relative maximal error on static bending<br />

<strong>MOE</strong>:<br />

Δ<strong>MOE</strong><br />

<strong>MOE</strong><br />

stat<br />

stat<br />

Δb<br />

Δh<br />

Δl<br />

s Δk<br />

= + 3 + 3 + (Equation 12)<br />

b h l k<br />

The relative errors on the measurements are based on a sample of 60 specimens <strong>and</strong> are:<br />

b = h = ± 1mm<br />

= ± 10 mm<br />

l s<br />

k is calculated using the 95% confidence interval from the slope of the <strong>for</strong>ce-deflection<br />

diagram. This method ensures the maximum relative error in the calculation of static <strong>MOE</strong><br />

using Equation 12.<br />

Δ k<br />

The coefficient is calculated as 3 % <strong>and</strong> due to the similar methodology used in this<br />

k<br />

study the same value was used. Thus the maximal relative error <strong>for</strong> static <strong>MOE</strong> is:<br />

Δ<strong>MOE</strong><br />

<strong>MOE</strong><br />

stat<br />

stat<br />

≈11%<br />

The maximal relative error of the <strong>MOR</strong> can be calculated by the following equation:<br />

s


Δ<strong>MOR</strong><br />

Δb<br />

Δh<br />

Δl<br />

s ΔP<br />

= + 2 + +<br />

<strong>MOR</strong> b h l s P<br />

The relative error of the load applied is known as 1 % <strong>for</strong> the machine used. There<strong>for</strong>e the<br />

maximal relative error of the modulus of rupture is:<br />

Δ<strong>MOR</strong><br />

<strong>MOR</strong><br />

≈ 6%<br />

Biased versus r<strong>and</strong>om position testing<br />

A biased test involves selection of a particular feature or position on the piece <strong>and</strong> deliberate<br />

placement in the centre of the test span in order to focus the results on the feature selected.<br />

For edge-biased bending tests, the selected feature is placed on the tension edge.<br />

A r<strong>and</strong>om test is where the central position of the test span is allocated using a r<strong>and</strong>om<br />

number generator to determine the measurement datum along the board without consideration<br />

of any properties or features of the timber in allocating test position.<br />

The issues of grade effectiveness must consider two opposing facts:<br />

• AS/NZS4063 is currently being revised, but will be underpinned by evaluation of<br />

characteristic values based on r<strong>and</strong>om-position testing. As a result, verification testing<br />

of stress-<strong>graded</strong> products will continue to be based on r<strong>and</strong>om-position tests.<br />

• The principle of stress-grading is founded on detecting the weakness in each piece to<br />

avoid failures in service where every part of most pieces will be used in some sort of<br />

role in buildings. The best grading methods will identify the lowest local strength of<br />

each piece rather than its global strength based on a r<strong>and</strong>om positioned test.<br />

The implications <strong>for</strong> grade evaluation are:<br />

• In per<strong>for</strong>ming normal commercial grading, the verification testing will normally use<br />

r<strong>and</strong>om-position tests <strong>and</strong> the test results will feed back to the threshold settings to<br />

ensure that the design characteristic values are reliably obtained by each <strong>graded</strong><br />

product.<br />

• In comparing grading methods, it is the correlation of grading parameter <strong>and</strong> the<br />

biased strength that is important. The most effective methods have higher correlation<br />

coefficients as they are most able to correctly predict the lowest local strength of each<br />

piece.<br />

For this study, most use will be made of the biased position test data. Successful grading<br />

methods will have a high correlation with the biased strength. Although it is possible to<br />

achieve meaningful results from r<strong>and</strong>om-position tests, very large sample sizes are required.<br />

The scatter of the results is generally high, as the test span may not contain any of the limiting<br />

material that actually determined the grade of the piece.<br />

Simulated MGP grades recovery by best predictors<br />

The static bending values obtained from the biased samples were used to find the best<br />

correlation <strong>for</strong> each non-destructive measurement system with the collected grading modulus.<br />

These correlations were identified by simple or multiple linear regressions.<br />

The grading modulus which provided the best correlations to the static bending data, were<br />

then used to allocate the boards into MGP grades. Thresholds were adjusted until the average<br />

of the <strong>MOE</strong> r<strong>and</strong>om <strong>and</strong> the 5 th percentile were equal or above the st<strong>and</strong>ard requirements. The<br />

average <strong>and</strong> 5 th percentile were calculated on the basis of the <strong>MOE</strong> <strong>and</strong> <strong>MOR</strong> obtained from<br />

26


the r<strong>and</strong>om sample tests. The outputs of various grading measures were used to group the test<br />

timber into notional grades <strong>and</strong> the test results used to confirm that the properties of each<br />

notional grade exceeded the design values.<br />

If the thresholds were adjusted <strong>for</strong> each group, one would optimise the recovery <strong>and</strong> the grade<br />

limiting properties would be much the same <strong>for</strong> each group, but the recovery would be<br />

different between the groups. This method would be very hard to implement in practice as a<br />

producer would have to know the optimum threshold to use be<strong>for</strong>e starting production, so that<br />

all of the pieces could be <strong>graded</strong> to give the grade limiting property just higher than the design<br />

value.<br />

A range of different thresholds was used <strong>and</strong> all of the results <strong>for</strong> each resource <strong>and</strong> <strong>for</strong> each<br />

prediction were plotted.<br />

Data extraction <strong>and</strong> statistical approach<br />

Automated data extraction was undertaken through collaboration with CIRAD Xylometry,<br />

Montpellier France. Mathematical <strong>and</strong> chemometrics’ statistical tools were used to provide<br />

correlations between data sets <strong>and</strong> to develop efficient strength <strong>and</strong> stiffness prediction<br />

equations. These algorithms were derived from raw <strong>and</strong> combined parameters using multivariate<br />

linear regression (MLR) <strong>and</strong> partial least squares (PLS).<br />

The following discussion on MLR <strong>and</strong> PLS was modified from statsoft.com (Web 2).<br />

For many data analysis problems, estimates of the linear relationships between variables are<br />

adequate to describe the observed data, <strong>and</strong> to make reasonable predictions <strong>for</strong> new<br />

observations. When the factors are few in number, are not significantly redundant (collinear),<br />

<strong>and</strong> have a well-understood relationship to the responses, then MLR can be a good way to<br />

turn data into in<strong>for</strong>mation.<br />

The MLR model serves as the basis <strong>for</strong> a number of multivariate methods such as:<br />

• discriminant analysis, i.e. the prediction of group membership from the levels of<br />

continuous predictor variables;<br />

• principal components regression, i.e. the prediction of responses on the dependent<br />

variables from factors underlying the levels of the predictor variables;<br />

• canonical correlation, i.e. the prediction of factors underlying responses on the<br />

dependent variables from factors underlying the levels of the predictor variables.<br />

These multivariate methods all have two important properties in common in that they impose<br />

restrictions such that:<br />

• factors underlying the Y <strong>and</strong> X variables are extracted from the Y'Y <strong>and</strong> X'X matrices,<br />

respectively, <strong>and</strong> never from cross-product matrices involving both the Y <strong>and</strong> X<br />

variables, <strong>and</strong><br />

• the number of prediction functions can never exceed the minimum of the number of Y<br />

variables <strong>and</strong> X variables.<br />

PLS has become a st<strong>and</strong>ard tool in chemometrics <strong>for</strong> modelling linear relations between<br />

multivariate measurements <strong>and</strong> is particularly suitable <strong>for</strong> monitoring <strong>and</strong> controlling<br />

industrial scenarios which may have hundreds of controllable variables <strong>and</strong> dozens of outputs.<br />

PLS represents a compromise between principal component regression (PCR) <strong>and</strong> MLR. PCR<br />

detects factors that account <strong>for</strong> the greatest amount of variance in the predictor variables.<br />

MLR seeks uncorrelated variables that best estimate the predicted variable. PLS screens <strong>for</strong><br />

factors which account <strong>for</strong> the variance <strong>and</strong> achieve the best correlation. A great advantage of<br />

27


PLS over other regression methods is that it can h<strong>and</strong>le large numbers of correlated variables<br />

as predictors. The method is based on projections whereby a set of correlated variables is<br />

compressed into a smaller set of uncorrelated variables.<br />

Partial least squares is a method <strong>for</strong> constructing predictive models when the factors are many<br />

<strong>and</strong> highly collinear. The emphasis is on predicting the responses <strong>and</strong> not necessarily on<br />

trying to underst<strong>and</strong> the underlying relationship between the variables. When prediction is the<br />

goal <strong>and</strong> there is no practical need to limit the number of measured factors, PLS can be a<br />

useful tool. Partial least squares regression is an extension of the multiple linear regression<br />

model (e.g., Multiple Regression or General Stepwise Regression). In its simplest <strong>for</strong>m, a<br />

linear model specifies the (linear) relationship between a dependent (response) variable Y,<br />

<strong>and</strong> a set of predictor variables, the X's, so that<br />

Y = b0 + b1X1 + b2X2 + ... + bpXp (Equation 13)<br />

In this equation b0 is the regression coefficient <strong>for</strong> the intercept <strong>and</strong> the bi values are the<br />

regression coefficients (<strong>for</strong> variables 1 through p) computed from the data.<br />

Partial least squares regression is probably the least restrictive of the various multivariate<br />

extensions of MLR. This flexibility allows it to be used in situations where the use of<br />

traditional multivariate methods is severely limited, such as when there are fewer<br />

observations than predictor variables. Furthermore, PLS regression can be used as an<br />

exploratory analysis tool to select suitable predictor variables <strong>and</strong> to identify outliers be<strong>for</strong>e<br />

classical linear regression.<br />

An important factor to consider when interpreting the results provided in this report is that the<br />

variation of the coefficient of determination is expressed as an absolute difference on a scale<br />

of 0.00 to 1.00, as opposed to a relative comparison.<br />

28


Results<br />

Static bending<br />

Biased <strong>and</strong> r<strong>and</strong>om tests together<br />

Table 1 displays basic statistics <strong>for</strong> dry density, <strong>MOE</strong> <strong>and</strong> <strong>MOR</strong> at 12% moisture content<br />

(MC) <strong>for</strong> all the boards tested (biased <strong>and</strong> r<strong>and</strong>om together) <strong>for</strong> the three resources. Caribbean<br />

has the highest density followed by Radiata R with Radiata E having the lowest density. The<br />

<strong>MOE</strong> follows logically the same trends, with a 5% difference between Radiata E <strong>and</strong> Radiata<br />

R <strong>and</strong> 21% between Caribbean <strong>and</strong> Radiata E. Caribbean <strong>MOR</strong> is 34% higher than Radiata R.<br />

The <strong>MOR</strong> <strong>for</strong> Radiata R is 3% lower than Radiata E <strong>MOR</strong> despite its higher <strong>MOE</strong>. The low<br />

Radiata R <strong>MOR</strong> confirms the resource quality observation by the West Australian industry<br />

which considers this resource as strength limited.<br />

The st<strong>and</strong>ard deviation of Caribbean density is significantly higher than the other two<br />

resources. <strong>MOE</strong> <strong>and</strong> <strong>MOR</strong> st<strong>and</strong>ard deviations of Radiata R <strong>and</strong> Caribbean are very close<br />

whereas Radiata E shows comparatively low st<strong>and</strong>ard deviations <strong>for</strong> both properties.<br />

Table 1. Descriptive statistics of density <strong>and</strong> mechanical properties <strong>for</strong> the three resources, biased <strong>and</strong><br />

r<strong>and</strong>om combined.<br />

Resource Mechanical properties Mean Std. Deviation N<br />

Radiata E Dry density (kg/m 3 ) 486 51 724<br />

<strong>MOE</strong> (MPa) 8317 2450 723<br />

<strong>MOR</strong> (MPa) 33 16 724<br />

Radiata R Dry density (kg/m 3 ) 508 41 689<br />

<strong>MOE</strong> (MPa) 8735 2911 686<br />

<strong>MOR</strong> (MPa) 32 20 690<br />

Caribbean Dry density (kg/m 3 ) 563 71 896<br />

<strong>MOE</strong> (MPa) 10048 2916 876<br />

<strong>MOR</strong> (MPa) 43 20 896<br />

The bivariate Pearson's correlation coefficients between density, <strong>MOE</strong> <strong>and</strong> <strong>MOR</strong> are<br />

displayed in Table 2 below.<br />

The <strong>MOE</strong> vs density correlations are quite different <strong>for</strong> each resource. Radiata E has the<br />

highest correlation coefficient followed by Radiata R with Caribbean having the lowest<br />

correlation coefficient. A similar trend is observed <strong>for</strong> <strong>MOE</strong> vs <strong>MOR</strong>.<br />

29


Table 2. Bivariate Pearson's correlation coefficients between density <strong>and</strong> mechanical properties <strong>for</strong> the<br />

three resources, biased <strong>and</strong> r<strong>and</strong>om combined.<br />

Resource Mechanical properties Dry density (kg/m 3 ) <strong>MOE</strong> (MPa)<br />

Radiata E <strong>MOE</strong> (MPa) 0.68<br />

<strong>MOR</strong> (MPa) 0.54 0.80<br />

Radiata R <strong>MOE</strong> (MPa) 0.54<br />

<strong>MOR</strong> (MPa) 0.33 0.77<br />

Caribbean <strong>MOE</strong> (MPa) 0.31<br />

<strong>MOR</strong> (MPa) 0.19 0.72<br />

Note: All the correlations are significant at the 0.01 level (2-tailed).<br />

Figure 8 shows the linear correlations between <strong>MOE</strong> <strong>and</strong> density <strong>for</strong> all the resources. From<br />

the scatter plot it appears that the Caribbean resource displays a significant number of outliers<br />

which have a high density with proportionally a low <strong>MOE</strong>. These outliers boards can be<br />

explained by the presence of a large quantity of resin <strong>and</strong> the occurrence of compression<br />

wood (high MFA) which both increase the density without significantly increasing the<br />

mechanical properties. Radiata E <strong>and</strong> Radiata R resources are different: the data shapes are<br />

not similar with Radiata E having a higher average <strong>MOE</strong> in the low density range.<br />

Figure 8. Linear regression between <strong>MOE</strong> <strong>and</strong> density <strong>for</strong> the three resources.<br />

The density vs <strong>MOR</strong> correlations show the same trend but exacerbated (Figure 9).<br />

30


Figure 9. Linear regression between <strong>MOR</strong> <strong>and</strong> density <strong>for</strong> the three resources.<br />

The <strong>MOR</strong> vs <strong>MOE</strong> linear correlations <strong>for</strong> all the resources are displayed in Figure 10. The<br />

general observation is the “trumpet” like shape of the data scatter points, in other words<br />

residual error on <strong>MOR</strong> increases with the <strong>MOE</strong>. This could be a source of heteroscedasticity<br />

problems (non-homogeneous variances) when developing linear regression equations.<br />

As expected the boards with the lowest strength belong to Radiata R resource. They are<br />

spread on a quite large span of <strong>MOE</strong>, up to approximately 12000 MPa. This induces a sort of<br />

“belly” on the Caribbean resource scatter plot. This is characteristic of a strength limited<br />

resource.<br />

The Caribbean resource has the lowest coefficient of determination which can be explained by<br />

a large variation of <strong>MOR</strong> values <strong>for</strong> a given upper range <strong>MOE</strong> (above 14000 MPa). Again the<br />

presence of a large quantity of resin associated with resin checks <strong>and</strong> the occurrence of<br />

compression wood could explain this observation.<br />

31


Figure 10. Linear regression between <strong>MOE</strong> <strong>and</strong> <strong>MOR</strong> <strong>for</strong> the three resources.<br />

R<strong>and</strong>om <strong>and</strong> biased<br />

Table 3 displays the three resources’ dry densities, <strong>MOE</strong>s <strong>and</strong> <strong>MOR</strong>s at 12% MC <strong>for</strong> all the<br />

boards tested by r<strong>and</strong>om <strong>and</strong> biased tests separately.<br />

The number of biased boards is larger than <strong>for</strong> r<strong>and</strong>om boards since the priority was given to<br />

biased position when per<strong>for</strong>ming the static bending test. The protocol was slightly different<br />

<strong>for</strong> Radiata E resource which explains the lower recovery of r<strong>and</strong>om boards (see Appendix 3),<br />

40% comparing to an average of 60% <strong>for</strong> the two other resources.<br />

The general tendencies observed when comparing the resources on all the boards taking all<br />

test positions together are similar to those observed on r<strong>and</strong>om or biased test positions. The<br />

mean density between biased <strong>and</strong> r<strong>and</strong>om is not significantly different whereas r<strong>and</strong>om <strong>and</strong><br />

biased <strong>MOE</strong> <strong>and</strong> <strong>MOR</strong> are obviously quite different.<br />

The total average <strong>MOE</strong> variations from biased to r<strong>and</strong>om relevant to the average <strong>MOE</strong> of all<br />

boards are 17%, 22% <strong>and</strong> 5% <strong>for</strong> Radiata E, Radiata R <strong>and</strong> Caribbean respectively. The total<br />

average <strong>MOR</strong> variations from biased to r<strong>and</strong>om relevant to the average <strong>MOR</strong> of all boards are<br />

54%, 59% <strong>and</strong> 32% <strong>for</strong> Radiata E, Radiata R <strong>and</strong> Caribbean respectively. The consequence of<br />

32


choosing biased or r<strong>and</strong>om has a smaller impact on Caribbean than it does on the two Radiata<br />

resources. The greater impact is on the Radiata R resource.<br />

Table 3. Descriptive statistics of density <strong>and</strong> mechanical properties <strong>for</strong> the three resources, biased <strong>and</strong><br />

r<strong>and</strong>om separated.<br />

Test position Resource Mean Std. Deviation N<br />

Biased Radiata E Dry density (kg/m 3 ) 484 51 517<br />

<strong>MOE</strong> (MPa) 7912 2349 516<br />

<strong>MOR</strong> (MPa) 28 13 517<br />

Radiata R Dry density (kg/m 3 ) 508 40 434<br />

<strong>MOE</strong> (MPa) 8041 2706 432<br />

<strong>MOR</strong> (MPa) 25 15 435<br />

Caribbean Dry density (kg/m 3 ) 564 71 558<br />

<strong>MOE</strong> (MPa) 9847 2842 541<br />

<strong>MOR</strong> (MPa) 38 17 558<br />

R<strong>and</strong>om Radiata E Dry density (kg/m 3 ) 489 51 207<br />

<strong>MOE</strong> (MPa) 9328 2411 207<br />

<strong>MOR</strong> (MPa) 46 17 207<br />

Radiata R Dry density (kg/m 3 ) 508 41 255<br />

<strong>MOE</strong> (MPa) 9915 2873 254<br />

<strong>MOR</strong> (MPa) 44 22 255<br />

Caribbean Dry density (kg/m 3 ) 561 72 338<br />

<strong>MOE</strong> (MPa) 10374 3007 335<br />

<strong>MOR</strong> (MPa) 52 21 338<br />

The histograms in Figure 11<strong>and</strong> Figure 12 illustrate the test consequence of the position<br />

choice <strong>for</strong> the three resources on <strong>MOE</strong> <strong>and</strong> <strong>MOR</strong>.<br />

33


Figure 11. Biased position <strong>MOE</strong> (left), <strong>and</strong> r<strong>and</strong>om position <strong>MOE</strong> (right), <strong>for</strong> all resources.<br />

Figure 12. Histograms <strong>for</strong> biased sample <strong>MOR</strong> (left), <strong>and</strong> r<strong>and</strong>om position <strong>MOR</strong> (right), all resources.<br />

The bivariate Pearson's correlation coefficients between density, <strong>MOE</strong> <strong>and</strong> <strong>MOR</strong> are<br />

displayed in Table 4 <strong>for</strong> biased <strong>and</strong> r<strong>and</strong>om tests separately. The correlations between <strong>MOE</strong><br />

34


<strong>and</strong> <strong>MOR</strong> with density are always better on r<strong>and</strong>om due to a wider span of values across the<br />

density range. Correlations between <strong>MOE</strong> <strong>and</strong> <strong>MOR</strong> are similar on biased <strong>and</strong> r<strong>and</strong>om <strong>for</strong><br />

Radiata E <strong>and</strong> R whereas Caribbean has better correlations on r<strong>and</strong>om.<br />

Table 4. Bivariate Pearson's correlation coefficients between density, <strong>MOE</strong> <strong>and</strong> <strong>MOR</strong>, biased <strong>and</strong><br />

r<strong>and</strong>om, all resources.<br />

Test<br />

Dry density<br />

Resource<br />

position<br />

(kg/m 3 <strong>MOE</strong><br />

) (MPa)<br />

Biased Radiata E <strong>MOE</strong> (MPa) 0.68<br />

<strong>MOR</strong> (MPa) 0.61 0.80<br />

Radiata R <strong>MOE</strong> (MPa) 0.52<br />

<strong>MOR</strong> (MPa) 0.34 0.74<br />

Caribbean <strong>MOE</strong> (MPa) 0.31<br />

<strong>MOR</strong> (MPa) 0.19 0.70<br />

R<strong>and</strong>om Radiata E <strong>MOE</strong> (MPa) 0.71<br />

<strong>MOR</strong> (MPa) 0.59 0.81<br />

Radiata R <strong>MOE</strong> (MPa) 0.65<br />

<strong>MOR</strong> (MPa) 0.41 0.75<br />

Caribbean <strong>MOE</strong> (MPa) 0.32<br />

<strong>MOR</strong> (MPa) 0.24 0.78<br />

All the correlations are significant at the 0.01 level (2-tailed).<br />

The latter observation on Caribbean can be explained by a larger dispersion of <strong>MOR</strong> values<br />

<strong>for</strong> <strong>MOE</strong> values above 12000 MPa (see Figure 13 <strong>and</strong> Figure 14).<br />

35


Figure 13. Biased test: <strong>MOE</strong> vs <strong>MOR</strong>, all resources.<br />

Figure 14. R<strong>and</strong>om test: <strong>MOE</strong> vs <strong>MOR</strong>, all resources.<br />

36


Biased vs R<strong>and</strong>om<br />

Figure 15, Figure 16 <strong>and</strong> Figure 17 display the correlations <strong>for</strong> each resource between <strong>MOR</strong><br />

r<strong>and</strong>om <strong>and</strong> <strong>MOR</strong> biased tests per<strong>for</strong>med on the same board at different locations. The red<br />

line represents the angle bisector of the graph. If selection of the weakest point on the board<br />

was perfect all the measurement points should be above this angle bisector line. For Radiata E<br />

resource the choice was generally good despite a few mistakes. In the case of Radiata R<br />

resource because some boards contained numerous large knots or defects it was sometimes<br />

difficult to nominate the weakest point unless without spending several minutes assessing<br />

each board. The result obtained <strong>for</strong> Caribbean is less easily explained because selection of the<br />

weakest point was apparently less difficult than Radiata R due to the proportion <strong>and</strong> sizes of<br />

the knots being somewhat smaller. Nevertheless it seems <strong>for</strong> some boards the nature or the<br />

type of the weakest knot or defect wasn’t assessed correctly be<strong>for</strong>e testing.<br />

37


Figure 15. Stiffness limited radiata pine, r<strong>and</strong>om vs biased <strong>MOR</strong>.<br />

Figure 16. Strength limited radiata pine, r<strong>and</strong>om vs biased <strong>MOR</strong>.<br />

38


Figure 17. Caribbean pine, r<strong>and</strong>om vs biased <strong>MOR</strong>.<br />

Position of the boards within the logs<br />

For all three resources the position of the board determined by distance from the pith within<br />

the log had only a weak correlation with the <strong>MOE</strong> or the <strong>MOR</strong> of the board when considering<br />

biased <strong>and</strong> r<strong>and</strong>om all together (Table 5). Caribbean displays the best correlation followed by<br />

Radiata E then Radiata R.<br />

Table 5. Distance from the pith vs static bending, all resources.<br />

Resource / distance from the pith - R 2<br />

<strong>MOE</strong> <strong>MOR</strong><br />

Radiata E 0.13 0.08<br />

Radiata R 0.09 0.04<br />

Caribbean 0.21 0.11<br />

39


Individual NDT predictors <strong>for</strong> static <strong>MOE</strong> <strong>and</strong> <strong>MOR</strong> biased evaluation<br />

In the following section all the analyses presented are based on biased boards only according<br />

to the justification given in the earlier section Biased versus r<strong>and</strong>om position testing.<br />

Correlation analyses between the individual NDT parameters <strong>and</strong> the destructively measured<br />

properties were carried out. In this way, the coefficient of determination (R 2 ) of each<br />

measured property in linear regression as a predictor <strong>for</strong> the destructively determined<br />

properties was obtained.<br />

Our logic in presenting the results has been to start from the data collected chronologically as<br />

close as possible to the reference static bending i.e. the dry boards then present the results in<br />

reverse chronological order viz the green boards <strong>and</strong> finally <strong>for</strong> the logs.<br />

At dry mill level<br />

The R 2 values <strong>for</strong> each parameter are derived from the dried, dressed boards. The indicating<br />

properties obtained from the strength grading machines or equipment used are included as<br />

single parameters, even if some of them required more than one measurement.<br />

Radiata E resource<br />

Figure 18 <strong>and</strong> Figure 19 present the coefficient of determination obtained on Radiata E<br />

resource between the different individual predictors <strong>and</strong> the biased static bending <strong>MOE</strong> <strong>and</strong><br />

<strong>MOR</strong> results.<br />

The best predictors <strong>for</strong> static <strong>MOE</strong> were achieved through the resonance method. We remind<br />

the reader that the dynamic <strong>MOE</strong> is based on the measurement combination of the vibration<br />

signal <strong>and</strong> the density. The flexion mode provides the higher coefficient of determination<br />

(Figure 18). Bernoulli’s model on the first resonance frequency <strong>and</strong> Timoshenko’s model both<br />

gave similar results. The second or third resonance frequencies with Bernoulli’s model gave a<br />

slightly lower level of correlation.<br />

The vibration <strong>MOE</strong> in compression mode is very close or even equivalent to the vibration<br />

<strong>MOE</strong> in flexion. The average <strong>MOE</strong> Metriguard profile on the reference destructive static test<br />

span (local measurement) provides a correlation 0.05 lower than the flexion resonance<br />

method. If the average Metriguard <strong>MOE</strong> profile is calculated on the full length of the board,<br />

the coefficient of determination is 0.04 less than the vibration <strong>MOE</strong> in compression. The<br />

minima <strong>MOE</strong> Metriguard provides a similar level of correlation to the Metriguard <strong>MOE</strong><br />

board profile. One would expect that the local measurement over the test span should be<br />

better than the global measurement (on the full length of the board), however this is not<br />

necessarily the case. It is possible that the flatwise, 3-point bending as utilised in the<br />

Metriguard process combined with the knotty attributes of the stiffness-limited resource,<br />

produces this discrepancy.<br />

Interestingly the specific vibration <strong>MOE</strong> in compression as a predictor of static bending <strong>MOE</strong><br />

provides an R 2 of 0.72 which is be acceptable to pre-sort the low stiffness boards with a very<br />

simple device able to measure only the board vibration. It may not be necessary to combine<br />

density as a parameter to achieve a satisfactory pre-grading result.<br />

The best WoodEye ® predictor <strong>for</strong> <strong>MOE</strong> is the KAR calculated with a triangular window<br />

(R 2 =0.31).<br />

40


Predictors <strong>MOE</strong> (MPa) <strong>MOR</strong> (MPa)<br />

<strong>MOE</strong> (MPa) 1.00 0.64<br />

Flex‐Dry‐<strong>MOE</strong>T<br />

0.85 0.53<br />

Flex‐Dry‐<strong>MOE</strong>_1<br />

0.85 0.53<br />

Comp‐Dry‐<strong>MOE</strong>_1<br />

0.83 0.50<br />

Flex‐Dry‐<strong>MOE</strong>_2<br />

0.82 0.50<br />

Flex‐Dry‐<strong>MOE</strong>_3<br />

0.82 0.50<br />

Metri_avg Static Test 0.80 0.46<br />

Comp‐Dry‐<strong>MOE</strong>_2<br />

0.80 0.48<br />

Flex‐Dry‐<strong>MOE</strong>_4<br />

0.80 0.48<br />

Metri_avg board 0.79 0.47<br />

Comp‐Dry‐<strong>MOE</strong>_3<br />

0.79 0.47<br />

Metri_min Static Test 0.78 0.50<br />

Metri_min board 0.77 0.52<br />

Flex‐Dry‐<strong>MOE</strong>_1 /Density<br />

0.75 0.41<br />

Comp‐Dry‐<strong>MOE</strong>_1/Density<br />

0.73 0.38<br />

Comp‐Dry‐<strong>MOE</strong>_4<br />

0.65 0.38<br />

<strong>MOR</strong> (MPa) 0.64 1.00<br />

Flex‐Dry‐SB<strong>and</strong>width<br />

0.64 0.35<br />

Flex‐Dry‐SCGravity<br />

0.53 0.28<br />

Comp‐Dry‐SB<strong>and</strong>width<br />

0.47 0.24<br />

Dry Density (kg/m3) 0.46 0.37<br />

Comp‐Dry‐SSlope<br />

0.32 0.19<br />

WE_KAR_trgl_99 0.31 0.29<br />

WE_KAR_ext_99 0.29 0.27<br />

WE_KAR_ext_98 0.28 0.27<br />

Comp‐Dry‐SCGravity<br />

0.27 0.13<br />

WE_CTR_trgl_AC_M 0.26 0.22<br />

WE_KAR_trgl_M 0.26 0.24<br />

WE_CTR_AC_M 0.26 0.22<br />

WE_KAR_ext_M 0.26 0.23<br />

WE_CTR_trgl_AC_3 0.24 0.24<br />

WE_CTR_trgl_AC_5 0.23 0.23<br />

Figure 18. R 2 of NDT measurements to destructively determined properties in bending <strong>for</strong> Radiata E<br />

resource on dry, dressed boards. Note: <strong>MOE</strong> R 2 sorted largest to smallest, lower limit = 0.2.<br />

The best prediction of <strong>MOR</strong> comes from static bending <strong>MOE</strong> which is appreciably above the<br />

second best predictor which is vibration <strong>MOE</strong> in flexion (Figure 18). Indeed the resonance<br />

method, whatever the configuration, <strong>and</strong> Metriguard give similar levels of prediction, around<br />

0.5. The <strong>MOE</strong>s calculated from higher resonance frequencies exhibit coefficient of<br />

determinations between 0.4 <strong>and</strong> 0.5. Because they are expressing the same physical property<br />

the levels of correlation with <strong>MOR</strong> are linked to those with static bending <strong>MOE</strong>. As a result,<br />

<strong>for</strong> simplification reasons, the Figure 18 displayed only the first resonance frequency <strong>MOE</strong>s<br />

in flexion <strong>and</strong> compression. Two signal descriptors, the spectral centre of gravity (R 2 =0.28)<br />

<strong>and</strong> the spectral b<strong>and</strong>width (R 2 =0.35), provide levels of prediction in the same range of the<br />

best optical scanner parameters.<br />

Dry density has an R 2 equal to 0.37, which is above the level of prediction of even the best<br />

WoodEye ® knot parameters which presents a coefficient of determination of 0.32. The fibre<br />

knot detection gives the best knot type prediction <strong>and</strong> the best mode of extraction is the<br />

maximum KAR.<br />

41


Predictors <strong>MOE</strong> (MPa) <strong>MOR</strong> (MPa)<br />

<strong>MOR</strong> (MPa) 0.64 1.00<br />

<strong>MOE</strong> (MPa) 1.00 0.64<br />

Flex‐Dry‐<strong>MOE</strong>_1 0.85 0.53<br />

Metri_min board 0.77 0.52<br />

Metri_min Static Test 0.78 0.50<br />

Comp‐Dry‐<strong>MOE</strong>_1 0.83 0.50<br />

Metri_avg board 0.79 0.47<br />

Metri_avg Static Test 0.80 0.46<br />

Dry Density (kg/m3) 0.46 0.37<br />

Fk_KAR_trgl_99<br />

Fk_KAR_ext_99<br />

Fk_KAR_ext_98<br />

WE_KAR_trgl_99<br />

WE_KAR_ext_99<br />

WE_KAR_ext_98<br />

Fk_CTR_trgl_AC_3<br />

Fk_CTR_trgl_AC_5<br />

Fk_KAR_trgl_M<br />

Fk_KAR_ext_M<br />

Fk_CTR_trgl_AC_M<br />

Fk_CTR_AC_M<br />

WE_CTR_trgl_AC_3<br />

WE_KAR_trgl_M<br />

WE_CTR_trgl_AC_5<br />

WE_KAR_ext_M<br />

WE_CTR_trgl_AC_M<br />

WE_CTR_AC_M<br />

0.33 0.32<br />

0.30 0.30<br />

0.30 0.30<br />

0.31 0.29<br />

0.29 0.27<br />

0.28 0.27<br />

0.26 0.27<br />

0.26 0.26<br />

0.27 0.25<br />

0.26 0.25<br />

0.29 0.25<br />

0.29 0.25<br />

0.24 0.24<br />

0.26 0.24<br />

0.23 0.23<br />

0.26 0.23<br />

0.26 0.22<br />

0.26 0.22<br />

Figure 19. R 2 of NDT measurements to destructively determined properties in bending <strong>for</strong> Radiata E<br />

resource on dry, dressed boards. Only the first frequency vibration <strong>MOE</strong> <strong>for</strong> both compression <strong>and</strong><br />

flexion are exhibited. Note: <strong>MOR</strong> R 2 sorted largest to smallest, lower limit = 0.2.<br />

Figure 20 displays the high correlation between Metriguard average <strong>MOE</strong> on the board <strong>and</strong><br />

vibration <strong>MOE</strong> in compression. The coefficients of determination are very high because the<br />

physical principles of measurement of these two methods are both established on strength of<br />

material principles <strong>and</strong> provide a direct measure of <strong>MOE</strong>. Nevertheless, they are based on<br />

different loading modes which explains the slight discrepancy observed through the linear<br />

regression equation (constant different from 0) <strong>and</strong> the residual error (dispersion around the<br />

regression line). Three factors can explain the observed difference of the experimental data:<br />

• Metriguard measures a long-span flat-wise average <strong>MOE</strong> whereas the vibration<br />

compression <strong>MOE</strong> is measured through a uni<strong>for</strong>m stress wave across the whole<br />

section.<br />

• The tested span is shorter <strong>for</strong> Metriguard because the measurement span starts <strong>and</strong><br />

finishes at 80 cm from both ends whereas the resonance method applies a stress on full<br />

length of the board.<br />

• The longitudinal resonance method induces a compression-traction stress whereas<br />

Metriguard system induces a three point bending stress. The latter provokes a shear<br />

effect which should lead to a lower measured <strong>MOE</strong>. One should note that the constant<br />

of the regression equation is the same order of magnitude as the shear modulus of<br />

elasticity in the longitudinal-radial plane of the wood. Indeed Metriguard provides a<br />

higher <strong>MOE</strong> than vibration <strong>MOE</strong> in compression. There is no obvious explanation to<br />

this phenomenon.<br />

42


Metriguard Average board (MPa)<br />

20000<br />

18000<br />

16000<br />

14000<br />

12000<br />

10000<br />

8000<br />

6000<br />

4000<br />

2000<br />

0<br />

y = x + 1590<br />

R² = 0.96<br />

0 2000 4000 6000 8000 10000 12000 14000 16000 18000 20000<br />

Vibration Compression <strong>MOE</strong>_1 (MPa)<br />

Figure 20. Linear regression between <strong>MOE</strong> measured by resonance method in compression <strong>and</strong><br />

average <strong>MOE</strong> given by Metriguard <strong>for</strong> Radiata E resource on dry, dressed boards.<br />

Radiata R resource<br />

Figure 21 <strong>and</strong> Figure 23 present the coefficient of determination obtained on Radiata R<br />

resource between the different individual predictors <strong>and</strong> the biased static bending <strong>MOE</strong> <strong>and</strong><br />

<strong>MOR</strong> results.<br />

For this resource the X-ray scanner LHG was also included in the set of methods evaluated.<br />

The Metriguard HCLT provided the strongest correlation with static <strong>MOE</strong>, using the<br />

minimum <strong>MOE</strong> from the profile on the static bending test span (its local measurement).<br />

The LHG predicted strength, which combines the previous parameter <strong>and</strong> X-ray knot<br />

descriptors, provided the second highest correlation as expected, due to the combination of<br />

Metriguard (minimum <strong>MOE</strong> from the profile on the static bending test span) <strong>and</strong> X-ray<br />

parameters. Vibration <strong>MOE</strong> in flexion, compression <strong>and</strong> flexion using Timoshenko’s model,<br />

all deliver comparable quality of prediction. These predictors are closely followed by the<br />

average static test (local measurement). Surprisingly, this local parameter (matching the test<br />

span <strong>for</strong> the reference test) did not provide a stronger correlation than the global<br />

measurements.<br />

Taken as a whole, the best predictors <strong>for</strong> the R limited resource have an average R 2 difference<br />

of 0.15 with Radiata E resource. This reflects the knotty nature of this resource which<br />

significantly impacts the results provided by individual or simple combined parameters.<br />

Specific knot in<strong>for</strong>mation should be qualitatively <strong>and</strong> quantitatively added to a prediction<br />

model in order to improve the predictions.<br />

As <strong>for</strong> Radiata E, the specific <strong>MOE</strong> correlation <strong>for</strong> the Radiata R was in the same magnitude<br />

of order to the best correlations, indicating that simple vibration systems may <strong>for</strong>m the basis<br />

of an effective pre-grading tool.<br />

43


For this resource, the R 2 <strong>for</strong> <strong>MOE</strong>s calculated using higher resonance frequencies were 0.05<br />

lower than the first frequency. The Metriguard average <strong>MOE</strong> displays R 2 significantly lower<br />

than other Metriguard <strong>and</strong> vibration parameters. Despite the characteristic knottiness of the<br />

Radiata R resource, the knot parameters as measured by WoodEye ® didn’t provide as good<br />

indications <strong>for</strong> stiffness as <strong>for</strong> the Radiata E resource.<br />

Predictors <strong>MOE</strong> (MPa) <strong>MOR</strong> (MPa)<br />

<strong>MOE</strong> (MPa) 1.00 0.56<br />

Metri_min Static Test 0.70 0.47<br />

LHG predicted Strength 0.67 0.53<br />

Flex‐Dry‐<strong>MOE</strong>_1<br />

0.66 0.28<br />

Comp‐Dry‐<strong>MOE</strong>_1<br />

0.65 0.28<br />

Flex‐Dry‐<strong>MOE</strong>T<br />

0.65 0.29<br />

Metri_avg Static Test 0.64 0.28<br />

Comp‐Dry‐<strong>MOE</strong>_1/Density<br />

0.63 0.26<br />

Metri_min board 0.61 0.37<br />

Flex‐Dry‐<strong>MOE</strong>_2<br />

Flex‐Dry‐<strong>MOE</strong>_3<br />

Flex‐Dry‐<strong>MOE</strong>_5<br />

Flex‐Dry‐<strong>MOE</strong>_4<br />

Flex‐Dry‐<strong>MOE</strong>_1 /Density<br />

Comp‐Dry‐<strong>MOE</strong>_2<br />

0.61 0.26<br />

0.61 0.26<br />

0.60 0.24<br />

0.59 0.23<br />

0.59 0.26<br />

0.57 0.25<br />

Metri_avg board 0.57 0.22<br />

<strong>MOR</strong> (MPa) 0.56 1.00<br />

Comp‐Dry‐<strong>MOE</strong>_3<br />

0.49 0.16<br />

Comp‐Dry‐<strong>MOE</strong>_4<br />

0.32 0.08<br />

Comp‐Dry‐SB<strong>and</strong>width<br />

0.32 0.16<br />

Comp‐Dry‐SSlope<br />

0.29 0.15<br />

Dry Density 0.28 0.12<br />

WE_KAR_trgl_M 0.27 0.30<br />

WE_KAR_ext_M 0.27 0.30<br />

WE_KAR_trgl_99 0.26 0.36<br />

WE_KAR_ext_99 0.24 0.35<br />

WE_KAR_ext_98 0.24 0.35<br />

Comp‐Dry‐SCGravity<br />

0.23 0.10<br />

Flex‐Dry‐SCGravity<br />

0.23 0.11<br />

Flex‐Dry‐SB<strong>and</strong>width<br />

0.23 0.12<br />

LHG KAR 0.21 0.25<br />

WE_CTR_trgl_AC_5 0.20 0.29<br />

Figure 21. R 2 of NDT-measurements to destructively determined properties in bending <strong>for</strong> Radiata R<br />

resource on dry, dressed boards. Note: <strong>MOE</strong> R 2 sorted largest to smallest, lower limit = 0.2.<br />

For <strong>MOR</strong> prediction, the <strong>MOE</strong>s calculated from higher resonance frequencies exhibit<br />

coefficient of determinations between 0.1 <strong>and</strong> 0.3. Because they are expressing the same<br />

physical property the levels of correlation with <strong>MOR</strong> are linked to those with static bending<br />

<strong>MOE</strong>. As a result, <strong>for</strong> simplification reasons, Figure 23 displays only the first resonance<br />

frequency <strong>MOE</strong>s in flexion <strong>and</strong> compression. Three signal descriptors, the spectral centre of<br />

gravity (R 2 =0.1), the spectral slope (R 2 =0.15) <strong>and</strong> the spectral b<strong>and</strong>width (R 2 =0.16), provide a<br />

level of prediction significantly lower than the best optical scanner parameters (R 2 between<br />

0.3 <strong>and</strong> 0.4).<br />

LHG predicted strength provided R 2 of 0.53, the best result from all <strong>MOR</strong> predictors (Figure<br />

23). As depicted in Figure 22 below, only the local measurements provided reasonable<br />

correlations <strong>for</strong> predictions, with Metriguard HCLT (minimum <strong>MOE</strong>) providing the best<br />

correlation to the static bending <strong>MOE</strong>. The test span in<strong>for</strong>mation provides a significantly<br />

superior prediction of strength (R 2 =0.47).<br />

44


The R 2 <strong>for</strong> Metriguard minimum of the board was 0.1 lower than the Metriguard Static test<br />

span result (0.37 vs 0.47 respectively). This implies that the minimum <strong>for</strong> the board is not the<br />

best predictor from the Metriguard <strong>MOE</strong> profile <strong>for</strong> this resource. In the case of the Radiata E<br />

resource, these two parameters provided a similar level of correlation.<br />

From the optical scanner in<strong>for</strong>mation, WoodEye ® KAR provided the best correlation with<br />

<strong>MOR</strong>, surpassing LHG X-ray by 0.1 (0.36 vs 0.25 respectively). The Metriguard minimum<br />

static test contains the majority of the in<strong>for</strong>mation <strong>and</strong> combining with LHG only improves by<br />

a magnitude of 0.06.<br />

Of the various knot types detected by WoodEye ® , fibre knots provide the best correlation,<br />

with a similar level of correlation to the combined knot silhouette.<br />

The Metriguard average <strong>and</strong> the vibration compression <strong>MOE</strong> both provide the same level of<br />

correlation.<br />

45


Predictors <strong>MOE</strong> (MPa) <strong>MOR</strong> (MPa)<br />

<strong>MOR</strong> (MPa) 0.56 1.00<br />

<strong>MOE</strong> (MPa) 1.00 0.56<br />

LHG predicted Strength 0.67 0.53<br />

Metri_min Static Test 0.70 0.47<br />

Metri_min board 0.61 0.37<br />

WE_KAR_trgl_99<br />

0.26 0.36<br />

WE_KAR_ext_99<br />

0.24 0.35<br />

WE_KAR_ext_98<br />

0.24 0.35<br />

WE_KAR_trgl_99<br />

0.25 0.34<br />

Fk_KAR_trgl_99<br />

0.23 0.34<br />

Fk_KAR_ext_99<br />

0.21 0.32<br />

Fk_KAR_ext_98<br />

0.21 0.32<br />

WE_KAR_ext_M<br />

0.27 0.30<br />

WE_KAR_trgl_M<br />

0.27 0.30<br />

Fk_CTR_trgl_AC_3<br />

0.21 0.30<br />

Fk_CTR_trgl_AC_5<br />

0.21 0.29<br />

WE_CTR_trgl_AC_5<br />

0.20 0.29<br />

WE_CTR_trgl_AC_3<br />

0.19 0.29<br />

Fk_KAR_trgl_M<br />

0.25 0.29<br />

Fk_KAR_ext_M<br />

0.24 0.28<br />

Flex‐Dry‐<strong>MOE</strong>_1 0.66 0.28<br />

Metri_avg Static Test 0.64 0.28<br />

Comp‐Dry‐<strong>MOE</strong>_1 0.65 0.28<br />

WE_KAR_trgl_M<br />

0.26 0.26<br />

LHG KAR 0.21 0.25<br />

LHG weighted KAR<br />

0.19 0.23<br />

Metri_avg board 0.57 0.22<br />

WE_KAR_int_99<br />

0.16 0.21<br />

WE_KAR_int_M<br />

0.18 0.19<br />

Fk_CTR_trgl_AC_M<br />

0.18 0.19<br />

Fk_CTR_AC_M<br />

0.18 0.19<br />

WE_CTR_trgl_AC_M<br />

0.18 0.19<br />

Fk_KAR_int_99<br />

0.13 0.18<br />

WE_CTR_AC_M<br />

0.18 0.18<br />

Fk_KAR_int_M<br />

0.16 0.18<br />

LHG I ratio 0.12 0.16<br />

Figure 23. R 2 of individual NDT-measurements to destructively determined properties in bending <strong>for</strong><br />

Radiata R resource on dried, dressed boards. Only the first frequency vibration <strong>MOE</strong> <strong>for</strong> both<br />

compression <strong>and</strong> flexion are exhibited. Note: <strong>MOR</strong> R 2 sorted largest to smallest, lower limit = LHG I<br />

ratio.<br />

Again we found a very good correlation <strong>for</strong> Metriguard average <strong>MOE</strong> with vibration<br />

compression <strong>MOE</strong> (R 2 = 0.90). However, due to the knot characteristics of the strength<br />

limited resource <strong>and</strong> subsequent outliers, as seen in Figure 24 below, the correlation is lower<br />

than was achieved <strong>for</strong> the stiffness limited resource.<br />

46


Metrigiard Average Board (MPa)<br />

20000<br />

18000<br />

16000<br />

14000<br />

12000<br />

10000<br />

8000<br />

6000<br />

4000<br />

2000<br />

0<br />

y = 0.9471x + 1912<br />

R² = 0.90<br />

0 2000 4000 6000 8000 10000 12000 14000 16000 18000 20000<br />

Vibration Compression <strong>MOE</strong>_1 (MPa)<br />

Figure 24. Linear regression between <strong>MOE</strong> measured by resonance method in compression <strong>and</strong><br />

average <strong>MOE</strong> given by Metriguard <strong>for</strong> Radiata R resource on dried, dressed boards.<br />

Caribbean resource<br />

The levels of correlation observed <strong>for</strong> the Caribbean biased static bending <strong>MOE</strong> are slightly<br />

lower than achieved <strong>for</strong> Radiata E. For this resource the Metriguard provided the best<br />

prediction of <strong>MOE</strong> (R 2 =0.84). As expected, the Metriguard average biased static test was<br />

better than the Metriguard average board result. In contrast to the Radiata E resource, the<br />

vibration <strong>MOE</strong> correlations were lower than <strong>for</strong> Metriguard.<br />

Vibration <strong>MOE</strong> (compression) without combining density still achieved a reasonable<br />

correlation (R 2 = 0.62). As <strong>for</strong> the other parameters, the vibration descriptors provided a range<br />

of R 2 from 0.30 to 0.40 <strong>for</strong> the best correlations.<br />

47


Predictors <strong>MOE</strong> (MPa) <strong>MOR</strong> (MPa)<br />

<strong>MOE</strong> (MPa) 1.00 0.49<br />

Metri_avg Static Test 0.84 0.35<br />

Metri_avg board 0.79 0.30<br />

Flex‐Dry‐<strong>MOE</strong>_3<br />

0.79 0.34<br />

Flex‐Dry‐<strong>MOE</strong>_4<br />

0.79 0.33<br />

Comp‐Dry‐<strong>MOE</strong>_1<br />

0.78 0.31<br />

Metri_min Static Test 0.77 0.37<br />

Flex‐Dry‐<strong>MOE</strong>T<br />

0.77 0.32<br />

Flex‐Dry‐<strong>MOE</strong>_5<br />

0.77 0.32<br />

Flex‐Dry‐<strong>MOE</strong>_2<br />

0.76 0.31<br />

Comp‐Dry‐<strong>MOE</strong>_2<br />

0.76 0.31<br />

Comp‐Dry‐<strong>MOE</strong>_3<br />

0.76 0.32<br />

Flex‐Dry‐<strong>MOE</strong>_1<br />

0.73 0.30<br />

Metri_min board 0.69 0.31<br />

Comp‐Dry‐<strong>MOE</strong>_1/Density<br />

0.62 0.23<br />

Comp‐Dry‐<strong>MOE</strong>_4<br />

0.53 0.23<br />

Flex‐Dry‐<strong>MOE</strong>_1 /Density<br />

0.50 0.20<br />

<strong>MOR</strong> (MPa) 0.49 1.00<br />

Comp‐Dry‐SSlope<br />

0.49 0.18<br />

Flex‐Dry‐SB<strong>and</strong>width<br />

0.46 0.18<br />

Flex‐Dry‐SCGravity<br />

0.43 0.16<br />

Comp‐Dry‐SB<strong>and</strong>width<br />

0.39 0.15<br />

Comp‐Dry‐SCGravity<br />

0.32 0.12<br />

Comp‐Dry‐MSNrjRatio_4<br />

0.24 0.09<br />

Comp‐Dry‐MSRatioF_3<br />

0.21 0.07<br />

Comp‐Dry‐MSPowerF_4<br />

0.20 0.10<br />

Figure 25. R 2 of individual NDT-measurements to destructively determined properties in bending <strong>for</strong><br />

Caribbean resource on dry, dressed boards. Note: <strong>MOE</strong> R 2 sorted largest to smallest, lower limit = 0.2.<br />

For <strong>MOR</strong> prediction, the <strong>MOE</strong>s calculated from higher resonance frequencies exhibit<br />

coefficient of determinations between 0.2 <strong>and</strong> 0.3. Because they are expressing the same<br />

physical property the levels of correlation with <strong>MOR</strong> are linked to those with static bending<br />

<strong>MOE</strong>. As a result, <strong>for</strong> simplification reasons, Figure 25 displays only the first resonance<br />

frequency <strong>MOE</strong>s in flexion <strong>and</strong> compression. Three signal descriptors, the spectral centre of<br />

gravity, the spectral slope <strong>and</strong> the spectral b<strong>and</strong>width, provided coefficient of determinations<br />

between 0.1 <strong>and</strong> 0.2. They are slightly lower than the best optical scanner parameters (R 2<br />

between 0.15 <strong>and</strong> 0.25).<br />

As <strong>for</strong> the <strong>MOE</strong> prediction, the best predictor of <strong>MOR</strong> <strong>for</strong> the Caribbean resource was<br />

provided by the Metriguard, local parameters (R 2 =0.37, Figure 26). This is significantly lower<br />

than <strong>for</strong> the Radiata E <strong>and</strong> Radiata R (both R 2 =0.53). Additionally, WoodEye ® provides a<br />

lower level of prediction in the case of Caribbean pine (R 2 = 0.24, best result <strong>for</strong> Caribbean<br />

compared to R 2 = 0.36 <strong>for</strong> Radiata R <strong>and</strong> R 2 =0.32 <strong>for</strong> Radiata E). The Caribbean pine knot<br />

attributes are generally captured by WoodEye ® scanners as ‘black knots’, whereas this feature<br />

is less frequently detected <strong>for</strong> the radiata resources, whose knot attributes are mostly detected<br />

as fibre knots. This difference in knot attributes/descriptors may explain the low level of<br />

correlation <strong>for</strong> <strong>MOR</strong>. This may be due to the incidence of loose knots in the Caribbean<br />

material.<br />

48


Predictors <strong>MOE</strong> (MPa) <strong>MOR</strong> (MPa)<br />

<strong>MOR</strong> (MPa) 0.49 1.00<br />

<strong>MOE</strong> (MPa) 1.00 0.49<br />

Metri_min Static Test 0.77 0.37<br />

Metri_avg Static Test 0.84 0.35<br />

Comp‐Dry‐<strong>MOE</strong>_1 0.78 0.31<br />

Metri_min board 0.69 0.31<br />

Flex‐Dry‐<strong>MOE</strong>_1 0.73 0.30<br />

Metri_avg board 0.79 0.30<br />

WE_KAR_ext_99<br />

WE_KAR_ext_98<br />

WE_KAR_trgl_99<br />

WE_CTR_trgl_AC_5<br />

WE_CTR_trgl_AC_3<br />

WE_KAR_ext_M<br />

WE_CTR_AC_5<br />

WE_KAR_trgl_M<br />

0.04 0.24<br />

0.04 0.24<br />

0.05 0.22<br />

0.04 0.20<br />

0.05 0.19<br />

0.03 0.18<br />

0.03 0.16<br />

0.03 0.16<br />

Bk_CTR_trgl_AC_3 0.06 0.16<br />

Fk_KAR_ext_99<br />

0.01 0.15<br />

Fk_KAR_ext_98<br />

0.01 0.15<br />

Bk_CTR_trgl_AC_5 0.05 0.15<br />

Bk_CTR_AC_5 0.05 0.14<br />

Fk_KAR_trgl_99<br />

0.01 0.14<br />

Fk_CTR_trgl_AC_5<br />

0.02 0.13<br />

Fk_CTR_trgl_AC_3<br />

0.02 0.13<br />

Bk_CTR_AC_M 0.05 0.12<br />

Bk_CTR_trgl_AC_M 0.05 0.12<br />

Fk_CTR_AC_5<br />

0.01 0.12<br />

Bk_KAR_trgl_99 0.02 0.11<br />

Fk_KAR_ext_M<br />

0.01 0.11<br />

WE_CTR_trgl_AC_M<br />

0.03 0.11<br />

Figure 26. R 2 of individual NDT-measurements to destructively determined properties in bending <strong>for</strong><br />

Caribbean resource on dry, dressed boards. Only the first frequency vibration <strong>MOE</strong> <strong>for</strong> both<br />

compression <strong>and</strong> flexion are exhibited. Note: <strong>MOR</strong> R 2 sorted largest to smallest, lower limit = 0.1.<br />

As was found <strong>for</strong> the two radiata resources, a high correlation was found between the<br />

Metriguard average board <strong>MOE</strong> <strong>and</strong> vibration compression <strong>MOE</strong> (R 2 =0.96, Figure 27).<br />

49


Metrigiard Average Board (MPa)<br />

25000<br />

20000<br />

15000<br />

10000<br />

5000<br />

0<br />

y = 0.95x + 1294<br />

R² = 0.96<br />

0 5000 10000 15000 20000 25000<br />

Vibration Compression <strong>MOE</strong>_1 (MPa)<br />

Figure 27. Linear regression between <strong>MOE</strong> measured by resonance method in compression <strong>and</strong><br />

average <strong>MOE</strong> given by Metriguard <strong>for</strong> Caribbean resource on dry, dressed boards.<br />

At green mill level<br />

Radiata E resource<br />

The vibration <strong>MOE</strong> in compression provides the best correlation with static bending <strong>MOE</strong><br />

<strong>and</strong> <strong>MOR</strong> (R 2 =0.77 <strong>and</strong> R 2 =0.47 respectively) as depicted in Figure 28. <strong>MOE</strong> from flexion<br />

vibration displays a slightly lower level of correlation (R 2 =0.75 <strong>and</strong> R 2 =0.45 respectively).<br />

The best predictor on dry dressed boards is from flexion vibration followed by compression<br />

vibration. They are calculated both from the first frequency with R 2 =0.85 <strong>and</strong> R 2 =0.83 <strong>for</strong><br />

<strong>MOE</strong> <strong>and</strong> R 2 =0.53 <strong>and</strong> R 2 =0.50 <strong>for</strong> <strong>MOR</strong> respectively. These are better than the green board<br />

prediction by a magnitude of 0.05 to 0.08. The specific compression vibration <strong>MOE</strong> as a<br />

single predictor didn’t provide a level of correlation suitable <strong>for</strong> sorting purposes.<br />

50


Predictors <strong>MOE</strong> (MPa) <strong>MOR</strong> (MPa)<br />

<strong>MOE</strong> (MPa) 1.00 0.64<br />

Comp‐Green‐<strong>MOE</strong>_1<br />

0.77 0.47<br />

Comp‐Green‐<strong>MOE</strong>_3<br />

0.76 0.45<br />

Flex‐Green‐<strong>MOE</strong>_3<br />

0.75 0.45<br />

Flex‐Green‐<strong>MOE</strong>T<br />

0.74 0.45<br />

Comp‐Green‐<strong>MOE</strong>_2<br />

0.74 0.45<br />

Flex‐Green‐<strong>MOE</strong>_2<br />

0.74 0.44<br />

Thumper_<strong>MOE</strong> (MPa) 0.74 0.44<br />

Flex‐Green‐<strong>MOE</strong>_1<br />

0.73 0.45<br />

Flex‐Green‐<strong>MOE</strong>_4<br />

0.72 0.44<br />

Comp‐Green‐<strong>MOE</strong>_4<br />

0.70 0.41<br />

<strong>MOR</strong> (MPa) 0.64 1.00<br />

Flex‐Green‐FacQ_4<br />

0.36 0.23<br />

Flex‐Green‐SSlope<br />

0.29 0.17<br />

Green Density 0.27 0.19<br />

Flex‐Green‐<strong>MOE</strong>_1 /Density<br />

0.26 0.13<br />

Comp‐Green‐<strong>MOE</strong>_1/Density<br />

0.25 0.12<br />

Gamma_Density(Kg/m3) 0.23 0.17<br />

Comp‐Green‐SSlope<br />

0.22 0.13<br />

Flex‐Green‐SB<strong>and</strong>width<br />

0.20 0.11<br />

Comp‐Green‐SB<strong>and</strong>width<br />

0.17 0.08<br />

Flex‐Green‐SCGravity<br />

0.13 0.07<br />

Flex‐Green‐R2T<br />

0.12 0.10<br />

Figure 28. R 2 of NDT measurements to destructively determined properties in bending <strong>for</strong> Radiata E<br />

resource on green sawn boards. Note: <strong>MOE</strong> R 2 sorted largest to smallest, lower limit = 0.1.<br />

This difference between green board <strong>and</strong> dry board prediction is statistically significant but it<br />

is still probably low enough to consider a pre-grading system <strong>for</strong> sorting the low stiffness<br />

green boards. This was partially done in one mill with the Thumper system which provides a<br />

high level of prediction equivalent to the one obtained with Bing® system. Indeed these two<br />

systems should provide the same <strong>MOE</strong> values.<br />

Figure 29 depicts the correlation between the green board compression <strong>MOE</strong> from the first<br />

frequency between Bing ® <strong>and</strong> Thumper system. The correlation is not as high as expected <strong>and</strong><br />

the residual error is quite significant. This discrepancy may be explained by the effect of<br />

uncontrolled drying on green density which occurred between the two set of measurements.<br />

Bing ® measurements were per<strong>for</strong>med three weeks after Thumper trial. The densities observed<br />

on the latter was on average 6% higher than the densities when the boards were measured <strong>for</strong><br />

Bing ® measurements with an important dispersion due to the position of the board in the stack<br />

(R 2 =0.86 between green density measured by direct weighing <strong>and</strong> gamma ray system).<br />

Moreover the acoustic equipment can differ, i.e. settings (support effect, dimensions<br />

measurements etc.) <strong>and</strong> extraction algorithms (peak detection) which can lead to some small<br />

differences in the <strong>MOE</strong> calculation.<br />

51


Thumper <strong>MOE</strong> (MPa)<br />

18000<br />

16000<br />

14000<br />

12000<br />

10000<br />

8000<br />

6000<br />

4000<br />

2000<br />

0<br />

y = 1.02x<br />

R² = 0.89<br />

0 2000 4000 6000 8000 10000 12000 14000 16000 18000<br />

Green compression Bing <strong>MOE</strong>_1 (MPa)<br />

Figure 29. Linear regression between <strong>MOE</strong> from resonance method in compression measured by off<br />

line Bing ® <strong>and</strong> by in-line Thumper systems, <strong>for</strong> Radiata E resource on green sawn boards.<br />

Radiata R resource<br />

Similarly to Radiata E resource, the vibration <strong>MOE</strong> in compression provides the best<br />

correlation with static bending <strong>MOE</strong> <strong>and</strong> <strong>MOR</strong> (R 2 =0.58 <strong>and</strong> R 2 =0.25 respectively) as<br />

depicted in Figure 30. <strong>MOE</strong> from flexion vibration displays a marginally lower level of<br />

correlation (R 2 =0.57 <strong>and</strong> R 2 =0.25 respectively).<br />

For this resource the level of correlation of the best predictors is around 0.2 lower than<br />

Radiata E. This difference is higher than the difference observed on dry boards (around 0.15).<br />

The drying process certainly plays a role in this difference since the differential shrinkage<br />

from the knotty area may induce cracks <strong>and</strong> lack of adherence which may contribute to<br />

blurring the correlations between green <strong>and</strong> dry products.<br />

Similar to Radiata E resource, the specific compression vibration <strong>MOE</strong> displays a very low<br />

prediction of static bending <strong>MOE</strong> <strong>and</strong> <strong>MOR</strong>.<br />

52


Predictors <strong>MOE</strong> (Mpa) <strong>MOR</strong> (Mpa)<br />

<strong>MOE</strong> (Mpa) 1.00 0.56<br />

Comp‐Green‐<strong>MOE</strong>_1<br />

0.58 0.25<br />

Flex‐Green‐<strong>MOE</strong>T<br />

0.57 0.25<br />

Flex‐Green‐<strong>MOE</strong>_1<br />

0.57 0.24<br />

<strong>MOR</strong> (Mpa) 0.56 1.00<br />

Flex‐Green‐<strong>MOE</strong>_2<br />

0.56 0.25<br />

Flex‐Green‐<strong>MOE</strong>_3<br />

0.54 0.22<br />

Comp‐Green‐<strong>MOE</strong>_2<br />

0.53 0.23<br />

Comp‐Green‐<strong>MOE</strong>_3<br />

0.53 0.23<br />

Flex‐Green‐<strong>MOE</strong>_5<br />

0.52 0.21<br />

Flex‐Green‐<strong>MOE</strong>_4<br />

0.51 0.21<br />

Comp‐Green‐<strong>MOE</strong>_4<br />

0.49 0.19<br />

Flex‐Green‐<strong>MOE</strong>_1/Density<br />

0.27 0.08<br />

Comp‐Green‐<strong>MOE</strong>_1/Density<br />

0.27 0.08<br />

Flex‐Green‐R2T<br />

0.27 0.15<br />

Comp‐Green‐SSlope<br />

0.19 0.09<br />

Flex‐Green‐SCGravity<br />

0.16 0.09<br />

Comp‐Green‐SB<strong>and</strong>width<br />

0.16 0.06<br />

Flex‐Green‐SB<strong>and</strong>width<br />

0.14 0.09<br />

Flex‐Green‐MSPowerF_2<br />

0.12 0.03<br />

Flex‐Green‐MSPower_1<br />

0.12 0.02<br />

Flex‐Green‐SSlope<br />

0.12 0.08<br />

Flex‐Green‐Q_2<br />

0.10 0.05<br />

Flex‐Green‐MSPowerF_3<br />

0.10 0.03<br />

Flex‐Green‐MSPower_2<br />

0.09 0.01<br />

Comp‐Green‐Q_4<br />

0.09 0.04<br />

Green Density 0.08 0.05<br />

Figure 30. R 2 of NDT measurements to destructively determined properties in bending <strong>for</strong> Radiata R<br />

resource on green sawn boards. Note: <strong>MOE</strong> R 2 sorted largest to smallest, lower limit given by the<br />

green density.<br />

Caribbean resource<br />

For the Caribbean resource, the best correlations are provided by flexion vibration which is<br />

marginally better than compression vibration. Contrasting with the other resources the best<br />

predictors are slightly improved from the use of higher resonance frequencies. The<br />

compression wood occurrence coupled together with high resin content may explain this<br />

observation already described in the static bending results paragraph above.<br />

The vibration <strong>MOE</strong> in flexion provides the best correlation with static bending <strong>MOE</strong> <strong>and</strong><br />

<strong>MOR</strong> (R 2 =0.75 <strong>and</strong> R 2 =0.31 respectively) as depicted in Figure 31. <strong>MOE</strong> from compression<br />

vibration displays a marginally lower level of correlation (R 2 =0.74 <strong>and</strong> R 2 =0.30 respectively).<br />

Apart from the vibration <strong>MOE</strong> extracted parameters, the signal descriptors provide a higher<br />

static bending <strong>MOE</strong> prediction than the other resources but not <strong>for</strong> <strong>MOR</strong> prediction.<br />

53


Predictors <strong>MOE</strong> (MPa) <strong>MOR</strong> (MPa)<br />

<strong>MOE</strong> (MPa) 1.00 0.49<br />

Flex‐Green‐<strong>MOE</strong>_4<br />

0.75 0.31<br />

Flex‐Green‐<strong>MOE</strong>_3<br />

0.75 0.32<br />

Flex‐Green‐<strong>MOE</strong>_2<br />

0.74 0.32<br />

Flex‐Green‐<strong>MOE</strong>T<br />

0.74 0.31<br />

Comp‐Green‐<strong>MOE</strong>_3<br />

0.74 0.30<br />

Comp‐Green‐<strong>MOE</strong>_2<br />

0.73 0.30<br />

Comp‐Green‐<strong>MOE</strong>_1<br />

0.73 0.30<br />

Comp‐Green‐<strong>MOE</strong>_4<br />

0.73 0.30<br />

Flex‐Green‐<strong>MOE</strong>_5<br />

0.73 0.30<br />

Flex‐Green‐<strong>MOE</strong>_1<br />

0.72 0.30<br />

Comp‐Green‐<strong>MOE</strong>_1/Density<br />

0.63 0.25<br />

Flex‐Green‐<strong>MOE</strong>_1/Density<br />

0.58 0.23<br />

<strong>MOR</strong> (MPa) 0.49 1.00<br />

Flex‐Green‐SB<strong>and</strong>width<br />

0.48 0.18<br />

Flex‐Green‐SSlope<br />

0.43 0.17<br />

Flex‐Green‐SCGravity<br />

0.43 0.16<br />

Comp‐Green‐SB<strong>and</strong>width<br />

0.41 0.16<br />

Comp‐Green‐SSlope<br />

0.37 0.13<br />

Comp‐Green‐SCGravity<br />

0.31 0.13<br />

Flex‐Green‐Q_2<br />

0.29 0.16<br />

Comp‐Green‐Alpha_1<br />

0.27 0.09<br />

Comp‐Green‐MSNrjRatio_3<br />

0.26 0.09<br />

Comp‐Green‐MSPower_3<br />

0.25 0.10<br />

Comp‐Green‐MSRatioF_3<br />

0.25 0.10<br />

Comp‐Green‐Q_2<br />

0.25 0.10<br />

Flex‐Green‐Alpha_2<br />

0.23 0.09<br />

Comp‐Green‐MSRatioF_4<br />

0.23 0.11<br />

Flex‐Green‐R2T<br />

0.22 0.09<br />

Comp‐Green‐MSNrjRatio_4<br />

0.22 0.10<br />

Green Density 0.22 0.09<br />

Figure 31. R 2 of NDT measurements to destructively determined properties in bending <strong>for</strong> Caribbean<br />

resource on green sawn boards. Note: <strong>MOE</strong> R 2 sorted largest to smallest, lower limit given by the<br />

green density.<br />

54


Combinations of NDT predictors <strong>for</strong> static <strong>MOE</strong> <strong>and</strong> <strong>MOR</strong> evaluation<br />

We remind the reader that only the results from the biased static bending tests are analyzed.<br />

The compression vibration system is more practical <strong>and</strong> easier to carry out in the mill than a<br />

flexion vibration because of the boundary conditions robustness <strong>and</strong> the fewer complementary<br />

data used. Moreover from the previous chapter we have noticed that the compression<br />

vibration <strong>and</strong> the flexion vibration provide comparable levels of prediction. For these reasons<br />

the following analysis will only focus on compression vibration data in combination with the<br />

other systems tested.<br />

In this chapter, two statistical approaches are used: the Multiple Linear Regression (MLR)<br />

<strong>and</strong> the Partial Least Squares (PLS).<br />

The variable selection <strong>for</strong> MLR was per<strong>for</strong>med through a stepwise process: the selection<br />

process starts by adding the variable with the largest contribution to the model (the criterion<br />

used is Student's t statistic). If a second variable is such that the probability associated with its<br />

t is less than the "Probability <strong>for</strong> entry", it is added to the model. The same protocol applies<br />

<strong>for</strong> a third variable. After the third variable is added, the impact of removing each variable<br />

present in the model after it has been added is evaluated (still using the t statistic). If the<br />

probability is greater than the "Probability of removal", the variable is removed. The<br />

procedure continues until no more variables can be added or removed. The collinearity (or<br />

multicollinearity) problem was prevented by limiting the number of variables introduced in<br />

the model through a condition index threshold. Heteroscedasticity (non-homogeneous<br />

variance) was checked. If the model was subject to an heteroscedasticity problem an<br />

appropriate data trans<strong>for</strong>mation was applied. This was often the case <strong>for</strong> <strong>MOR</strong> prediction<br />

where a logarithm-trans<strong>for</strong>m was per<strong>for</strong>med.<br />

PLS attempts to find factors which both capture variance <strong>and</strong> achieve correlation. PLS<br />

attempts to maximize covariance <strong>and</strong> this is the explicit objective of the simple PLS<br />

(SIMPLS) algorithm used in this study. Two rules of thumb with regard to selecting the<br />

optimal number of factors were applied, namely:<br />

• to only choose additional factors when the RMSEC improves by at least 2%, <strong>and</strong><br />

• to choose as few factors (called Latent Variables) as possible.<br />

When developing the models an auto-scaling process was used (mean-centring followed by<br />

dividing variable by the st<strong>and</strong>ard) <strong>and</strong> no cross-validation per<strong>for</strong>med. A log trans<strong>for</strong>mation<br />

was systematically applied on <strong>MOR</strong> values.<br />

Radiata E resource<br />

The coefficient of determination <strong>for</strong> predicting the static bending <strong>MOE</strong> from the best<br />

combinations of methods or parameters on dry boards ranges from 0.79 to 0.86 (Table 6).<br />

Using average <strong>MOE</strong> <strong>and</strong> minimum <strong>MOE</strong> together from Metriguard profiles in MLR doesn’t<br />

improve the correlation. Only the profile average provides the best prediction from this<br />

equipment on Radiata E resource.<br />

From the 35 variables extracted from each vibration spectrum only the first frequency<br />

vibration <strong>MOE</strong> was selected as a suitable explanatory variable in MLR. This fact is confirmed<br />

by the PLS model which displays the same level of prediction despite its complexity. The<br />

55


only advantage to use a PLS approach could be the robustness of the model since it uses<br />

several variables to per<strong>for</strong>m the prediction.<br />

Adding WoodEye ® <strong>and</strong>/or Metriguard in<strong>for</strong>mation to vibration measurement only improved<br />

the level of prediction by a maximum of 0.03. This observation indicates that the local<br />

in<strong>for</strong>mation provided by WoodEye ® or Metriguard marginally improves the static bending<br />

<strong>MOE</strong> prediction.<br />

For the green boards, the static bending <strong>MOE</strong> prediction level is below that observed <strong>for</strong> dry<br />

boards by an average of only 0.06. If an optical or equivalent system can be used on green<br />

boards with the same quality of in<strong>for</strong>mation as provided on dry boards, then there is a small<br />

improvement when combined with vibration using PLS on all parameters (+0.02).<br />

Table 6. R 2 of combined NDT measurements to destructively determined properties in bending <strong>for</strong><br />

Radiata E resource on green sawn boards. Note that the number listed beside PLS indicates the<br />

number of factors used in the models. Each colour corresponds to a different set of equipment or wood<br />

moisture content.<br />

Methods Reg. Parameters R2 x 100<br />

<strong>MOE</strong><br />

Vibration GREEN<br />

Vibration GREEN +<br />

WoodEye<br />

MLR/P<br />

LS 3<br />

56<br />

Parameters<br />

R 2 x 100<br />

<strong>MOR</strong><br />

<strong>MOE</strong>_1 77 <strong>MOE</strong>_1 48<br />

PLS 3 all 79 all 55<br />

Metriguard Board MLR Avg_Brd 79 Mini_Brd 52<br />

Metriguard Test Span MLR Avg_Span 80 Mini_Span 52<br />

Vibration MLR <strong>MOE</strong>_1 83<br />

<strong>MOE</strong>_1 +<br />

<strong>MOE</strong>_1/density +<br />

Alpha_2<br />

Vibration PLS 4 all 83 all 57<br />

Vibration + Metriguard MLR<br />

<strong>MOE</strong>_1 +<br />

Mini_Span<br />

53<br />

85 Mini_Span + <strong>MOE</strong>_1 54<br />

Vibration + Metriguard PLS 4 all 85 all 61<br />

Vibration + WoodEye MLR<br />

<strong>MOE</strong>_1 +<br />

Fk_KAR-m<br />

84 <strong>MOE</strong>_1 + KAR-Ext 55<br />

Vibration + WoodEye PLS 3 all 84 all 62<br />

Vibration + WoodEye +<br />

Metriguard<br />

Vibration + WoodEye +<br />

Metriguard<br />

MLR<br />

<strong>MOE</strong>_1 +<br />

Mini-Span<br />

85<br />

Mini_Brd + Kar-Ext +<br />

<strong>MOE</strong>_1<br />

PLS 3 all 86 all (PLS 4) 64<br />

When considering the <strong>MOR</strong> predictions, MLR provided an R 2 of 0.52 up to 0.57, with the<br />

best combination being the combination of three systems: vibration, WoodEye ® <strong>and</strong><br />

Metriguard. This level of prediction is achieved by using only the vibration methods in<br />

combination with a PLS model. In combination with Metriguard or WoodEye ® , the prediction<br />

increases to 0.62. The best combination is achieved with the three systems combined where<br />

an R 2 of 0.64 resulted through PLS approach.<br />

57


The first frequency vibration <strong>MOE</strong> alone provided an R 2 of 0.48 which wasn’t improved by<br />

using MLR or PLS. Interestingly, using PLS on vibration parameters <strong>for</strong> green boards<br />

combined with an optical scanner achieved a prediction level as good as MLR on dry boards<br />

using a combination of vibration <strong>and</strong> Metriguard or WoodEye ® (R 2 =0.55).<br />

Radiata R resource<br />

On this resource, Metriguard only achieved an R 2 of 0.70 <strong>for</strong> static bending <strong>MOE</strong> prediction,<br />

which is 0.1 below the other two resources. The compression vibration system, whatever the<br />

statistical method, doesn’t provide a better correlation; however, combining vibration<br />

measurements with local in<strong>for</strong>mation obtained from WoodEye ® or Metriguard, significantly<br />

improves the prediction (up to 0.77). As shown in Table 7 there is only marginal<br />

improvement by combining all three methods.<br />

Combining LHG with vibration doesn’t significantly improve the <strong>MOE</strong> prediction when<br />

compared to Metriguard alone. Metriguard combined with LHG provided an R 2 between<br />

Metriguard <strong>and</strong> vibration plus Metriguard or WoodEye ® . The use of PLS method of<br />

regression doesn’t improve <strong>MOE</strong> prediction compared to using MLR using 2 or 3 parameters.<br />

On green boards when using PLS method, vibration alone indicates a level of correlation<br />

equivalent to the Metriguard average <strong>MOE</strong> profile on the board. If an optical or equivalent<br />

system can be used on green boards with the same quality of in<strong>for</strong>mation as provided on dry<br />

boards, then there is a large improvement when combined with vibration using PLS on all<br />

parameters (approximately +0.1).<br />

Contrary to Radiata E, these results prove that the characteristic knotty nature of his resource<br />

heavily impacts the <strong>MOE</strong>, since when local parameters are combined with global<br />

measurements the prediction improvement is substantial.<br />

57


Table 7. R 2 of combined NDT measurements to destructively determined properties in bending <strong>for</strong><br />

Radiata R resource on green sawn boards. Note that the number listed beside PLS indicates the<br />

number of factors used in the models. Each colour corresponds to a different set of equipment or wood<br />

moisture content.<br />

Methods Reg. Parameters<br />

R 2 x 100<br />

<strong>MOE</strong><br />

Parameters<br />

R 2 x 100<br />

<strong>MOR</strong><br />

Vibration GREEN MLR <strong>MOE</strong>_1 59 <strong>MOE</strong>_1 26<br />

Vibration GREEN PLS 4 all 62 all 33<br />

Vibration GREEN +<br />

WoodEye<br />

PLS3 all 71 all 55<br />

Metriguard Board MLR Avg_Brd 63 Mini_Brd 27<br />

Metriguard Test span MLR Avg_Span 70 Mini_Span 47<br />

Vibration MLR<br />

<strong>MOE</strong>_1 +<br />

Mspower<br />

58<br />

67<br />

<strong>MOE</strong>_1 +<br />

<strong>MOE</strong>_1/density +<br />

Alpha_2<br />

Vibration PLS 4 all 68 all 38<br />

Vibration + Metriguard MLR<br />

<strong>MOE</strong>_1 +<br />

Mini_Span<br />

77<br />

Mini_Span +<br />

MSpower2<br />

48<br />

Vibration + Metriguard PLS 4 all 77 all 49<br />

Vibration + WoodEye MLR<br />

<strong>MOE</strong>_1 +<br />

KAR_Ext<br />

75 <strong>MOE</strong>_1 + KAR_Ext 49<br />

Vibration + WoodEye PLS 4 all 76 all 59<br />

Vibration + WoodEye +<br />

Metriguard<br />

MLR<br />

<strong>MOE</strong>_1 +<br />

KAR_Ext +<br />

Mini_Span<br />

78<br />

<strong>MOE</strong>_1 + KAR_Ext<br />

+ MSPower2<br />

57<br />

Vibration + WoodEye +<br />

Metriguard<br />

PLS 4 all 79 all 60<br />

Vibration + LHG<br />

(without Strength<br />

prediction)<br />

MLR<br />

<strong>MOE</strong>_1 +<br />

Weighted_KAR<br />

70<br />

<strong>MOE</strong>_1 +<br />

Weighted_KAR +<br />

MSpower2+ LHG Iratio<br />

43<br />

LHG strength prediction<br />

Vibration + LHG<br />

LR<br />

Indicative<br />

property<br />

67 Indicative property 53<br />

(without Strength<br />

prediction)<br />

PLS 4 all 71 all 48<br />

Metriguard + LHG MLR<br />

Mini_Span +<br />

Avg_brd<br />

73<br />

Mini_Span + LHG I<br />

Ratio<br />

51<br />

Metriguard + LHG PLS<br />

not enough<br />

variables<br />

- not enough variables -<br />

The level of prediction <strong>for</strong> strength is rather low <strong>for</strong> most methods. The trend is the same as<br />

<strong>for</strong> <strong>MOE</strong>, in that when local in<strong>for</strong>mation <strong>and</strong> global in<strong>for</strong>mation are combined the prediction<br />

is significantly improved. Interestingly the PLS regression using data from vibration <strong>and</strong><br />

WoodEye ® systems provides one of the highest prediction of strength (R 2 =0.59). The R 2 <strong>for</strong><br />

LHG strength prediction is 0.53 which is the parameter currently used by the processor with<br />

this equipment. With additional optical equipment this may be improved significantly.<br />

Caribbean resource<br />

Similarly to Radiata E resource, the coefficient of determination <strong>for</strong> predicting the static<br />

bending <strong>MOE</strong> from the best combinations of methods or parameters on dry boards, ranges<br />

from 0.79 to 0.87 (Table 8). Using average <strong>MOE</strong> from Metriguard profiles from the static<br />

31


ending span test significantly improves the correlation compared to using the average of the<br />

board.<br />

Vibration system alone provides a level of prediction equivalent to Metriguard from the<br />

average <strong>MOE</strong> on the board. These two systems combine together doesn’t improve the static<br />

bending <strong>MOE</strong>. The average <strong>MOE</strong> from Metriguard profiles on static bending span remains<br />

the only reliable predictor since the MLR stepwise process selects only this parameter to<br />

per<strong>for</strong>m the prediction.<br />

Adding WoodEye ® <strong>and</strong>/or Metriguard in<strong>for</strong>mation to vibration measurement only improves<br />

the level of prediction by a maximum of 0.05. This observation indicates that the local<br />

in<strong>for</strong>mation provided by WoodEye ® or Metriguard improves the static bending <strong>MOE</strong><br />

prediction. The best prediction (R 2 =0.87) was achieved by using the WoodEye ® (KAR outer<br />

third zone) <strong>and</strong> Metriguard (average <strong>MOE</strong> on test span) combined, without any measurable<br />

improvement by adding vibration to the combination. The best predictions are close to the<br />

maximum feasible level of correlation <strong>for</strong> <strong>MOE</strong>. Indeed the coefficient of determination<br />

reaches an asymptote thus it is very hard to improve the correlation as the measurement error<br />

is the main limiting factor.<br />

For the green boards, the static bending <strong>MOE</strong> prediction level is below that observed <strong>for</strong> dry<br />

boards by an average of 0.08. If an optical or equivalent system can be used on green boards<br />

with the same quality of in<strong>for</strong>mation as provided on dry boards, then there is a small<br />

improvement when combined with vibration using PLS on all parameters (+0.02).<br />

Table 8. R 2 of combined NDT measurements to destructively determined properties in bending <strong>for</strong><br />

Caribbean resource on green sawn boards. Note: <strong>MOE</strong> R 2 sorted largest to smallest, lower limit given<br />

by the green density. Note that the number listed beside PLS indicates the number of factors used in<br />

the models. Each colour corresponds to a different set of equipment or wood moisture content.<br />

Methods Reg. Parameters<br />

R 2 x 100<br />

<strong>MOE</strong><br />

Parameters<br />

R 2 x 100<br />

<strong>MOR</strong><br />

Vibration GREEN MLR <strong>MOE</strong>_1 73 <strong>MOE</strong>_1 31<br />

Vibration GREEN PLS 3 all 75 all 34<br />

Vibration GREEN +<br />

WoodEye<br />

PLS 2 all 77 all (PLS 3) 55<br />

Metriguard Board MLR Avg_Brd 79 Mini_Brd 30<br />

Metriguard Test Span MLR Avg_Span 84 Mini_Span 36<br />

Vibration MLR <strong>MOE</strong>_1 + FacQ1 78 <strong>MOE</strong>_1 32<br />

Vibration PLS 4 all 79 all 33<br />

Vibration + Metriguard MLR Avg_Span 84 Mini_Span 36<br />

Vibration + Metriguard PLS 3 all 82 all 36<br />

Vibration + WoodEye MLR<br />

<strong>MOE</strong>_1 +<br />

Fk_KAR-Ext<br />

79<br />

<strong>MOE</strong>_2 +<br />

KAR-Ext<br />

46<br />

Vibration + WoodEye PLS 4 all 84 All (PLS 3) 55<br />

Vibration + WoodEye +<br />

Metriguard<br />

MLR<br />

Avg_Span + KAR-<br />

Ext<br />

87<br />

Mini_Span+<br />

Kar-Ext<br />

49<br />

Vibration + WoodEye +<br />

Metriguard<br />

PLS 3 all 83 all 56<br />

Despite the Caribbean resource having a similar range in the results <strong>for</strong> <strong>MOE</strong> prediction as <strong>for</strong><br />

the Radiata E resource (R 2 =0.79 to 0.87 <strong>and</strong> 0.79 to 0.86 respectively), the level of accuracy<br />

was not as strong <strong>for</strong> strength prediction (R 2= 0.30 to 0.55 <strong>and</strong> 0.52 to 0.64). The poor level of<br />

59


strength prediction <strong>for</strong> Caribbean pine using current systems may be attributable to a feature<br />

of the knots in the resource tested, as discussed in the earlier section on individual NDT<br />

predictors.<br />

The same level of strength prediction achieved on dry boards was achieved on green boards<br />

(R 2 =0.55). This calculation was determined by the combined vibration (green boards) <strong>and</strong><br />

WoodEye ® (dry boards) data, <strong>and</strong> assumes that the optical data can be captured at the green<br />

stage with equivalent accuracy to that captured on dry boards. The use of PLS regression<br />

provides the best predictions.<br />

Prediction of static bending <strong>MOE</strong> <strong>and</strong> <strong>MOR</strong> from logs<br />

This section uses the average properties of all the boards from the log as the basis <strong>for</strong> the<br />

correlations. It does not use log property as a predictor of individual board properties. A<br />

summary of results is provided in Table 9.<br />

Radiata E<br />

The <strong>MOE</strong> velocity was calculated using Equation 1, combination of log density data <strong>and</strong><br />

velocity (TOF) measurements in the longitudinal axes. When correlated with the average<br />

static <strong>MOE</strong> levels of prediction of R 2 of 0.64 <strong>for</strong> <strong>MOE</strong> <strong>and</strong> 0.33 <strong>for</strong> <strong>MOR</strong> were provided.<br />

Slightly better correlations were observed when using velocity alone (TOF <strong>and</strong> distance<br />

between sensors) to predict <strong>MOE</strong> <strong>and</strong> <strong>MOR</strong>. These are lower than the prediction provided by<br />

vibration analyses (R 2 =0.70 <strong>and</strong> 0.45 respectively).<br />

The MLR combining compression <strong>and</strong> flexion tests increases the coefficient of determination<br />

from 0.70 to 0.76 <strong>for</strong> <strong>MOE</strong> <strong>and</strong> 0.45 to 0.61 <strong>for</strong> <strong>MOR</strong>, significant improvements, especially in<br />

the case of <strong>MOR</strong> prediction. However, this combination of measurements would require<br />

additional systems to enable measurements in the two planes. When developing the stepwise<br />

MLR to combine compression <strong>and</strong> flexion parameters, the condition index was close to the<br />

threshold, indicating a potential collinearity problem. The quality of this prediction needs to<br />

be considered with care until a larger population of logs can be tested <strong>for</strong> verification.<br />

Radiata R<br />

The velocity <strong>MOE</strong> measurements provided an R 2 of 0.50 <strong>for</strong> static <strong>MOE</strong> <strong>and</strong> 0.34 <strong>for</strong> <strong>MOR</strong>.<br />

The <strong>MOE</strong> prediction from logs <strong>for</strong> this resource is quite significantly lower compared to the<br />

Radiata E prediction (-0.14). The <strong>MOR</strong> predictions were similar <strong>for</strong> both resources.<br />

Noticeably better correlations were observed when using velocity alone (TOF <strong>and</strong> distance<br />

between sensors) to predict <strong>MOE</strong> <strong>and</strong> <strong>MOR</strong>.<br />

Vibration analysis provides better predictions compared to velocity measurements, with the<br />

best result using MLR with specific compression <strong>MOE</strong> (R 2 =0.75). When predicting strength<br />

using the same parameter from the log tests, the R 2 achieved was 0.43. These results are<br />

interesting due to the fact that only the first frequency measurement was required <strong>and</strong><br />

there<strong>for</strong>e the density data wasn’t necessary to improve the result. This has positive<br />

implications <strong>for</strong> in-mill engineering <strong>and</strong> logistics in that weighing <strong>and</strong> log measuring systems<br />

may not be necessary <strong>for</strong> <strong>MOE</strong> prediction, but a larger log population should be tested to<br />

verify these results.<br />

60


Caribbean<br />

The velocity <strong>MOE</strong> measurements provided an R 2 of 0.56 <strong>for</strong> static <strong>MOE</strong> <strong>and</strong> 0.27 <strong>for</strong> <strong>MOR</strong>.<br />

The <strong>MOE</strong> prediction from logs <strong>for</strong> this resource is between the two radiata resources. The<br />

<strong>MOR</strong> prediction was significantly lower than the other material (-0.06). Conversely to the two<br />

other resources, markedly poorer correlations were observed when using velocity alone (TOF<br />

<strong>and</strong> distance between sensors) to predict <strong>MOE</strong> <strong>and</strong> <strong>MOR</strong>.<br />

As with radiata pine, vibration analyses provided better predictions compared to velocity<br />

measurements.<br />

Table 9. R 2 of combined NDT measurements to log average, destructively determined properties in<br />

bending <strong>for</strong> all resources. Note that the number listed beside PLS indicates the number of factors used<br />

in the models. Each colour corresponds to a different set of measurement.<br />

Resource<br />

Biased +<br />

R<strong>and</strong>om<br />

Methods Reg.<br />

type<br />

Parameters<br />

61<br />

R 2 x 100<br />

<strong>MOE</strong><br />

Parameters<br />

R 2 x<br />

100<br />

<strong>MOR</strong><br />

Velocity<br />

(TOF)<br />

<strong>MOE</strong><br />

LR Velocity 65 Velocity 36<br />

Velocity<br />

(TOF)<br />

LR Velocity_<strong>MOE</strong> 64 Velocity_<strong>MOE</strong> 33<br />

Radiata E<br />

(n= 34)<br />

Specific<br />

<strong>MOE</strong><br />

LR<br />

Comp-<strong>MOE</strong>_1<br />

/density<br />

65 Comp-<strong>MOE</strong>_1 /density 43<br />

Comp LR Comp_<strong>MOE</strong>_1 70 comp-SB<strong>and</strong>width 45<br />

Comp +<br />

Flex<br />

MLR<br />

Comp_<strong>MOE</strong>_1,<br />

Ratio (Comp/Flex)<br />

76<br />

comp-<strong>MOE</strong>_1, Flex-<br />

MSPower_0<br />

61<br />

Velocity<br />

(TOF)<br />

<strong>MOE</strong><br />

LR Velocity 65 Velocity 46<br />

Velocity LR Velocity_<strong>MOE</strong> 50 Velocity_<strong>MOE</strong> 34<br />

Radiata R<br />

(n= 27)<br />

(TOF)<br />

Specific<br />

<strong>MOE</strong><br />

LR<br />

Comp-<strong>MOE</strong>_1<br />

/density<br />

75 Comp-<strong>MOE</strong>_1 /density 43<br />

Comp MLR<br />

Comp-<strong>MOE</strong>_1<br />

/density<br />

75 Comp-<strong>MOE</strong>_1 /density 43<br />

Velocity<br />

(TOF)<br />

<strong>MOE</strong><br />

LR Velocity 48 Velocity 22<br />

Velocity<br />

(TOF)<br />

LR Velocity_<strong>MOE</strong> 56 Velocity_<strong>MOE</strong> 27<br />

Caribbean<br />

(n= 106)<br />

Specific<br />

<strong>MOE</strong><br />

LR<br />

Comp-<strong>MOE</strong>_1<br />

/density<br />

59 Comp-<strong>MOE</strong>_1 /density 27<br />

Comp MLR<br />

Comp-<strong>MOE</strong>_4,<br />

Comp-Alpha_2<br />

75<br />

Comp-<strong>MOE</strong>_3, Comp-<br />

Alpha_2<br />

37<br />

Comp PLS 3 all 78 all 44<br />

PLS analysis was able to be used on the Caribbean resource due to the higher number of logs<br />

in the study. As with green boards, the improvement <strong>for</strong> <strong>MOE</strong> prediction using the PLS<br />

method <strong>for</strong> Caribbean logs is slight, but <strong>for</strong> <strong>MOR</strong> it is a significant improvement.<br />

The reader should note that when using the data <strong>for</strong> all the boards per log, instead of the<br />

average of all boards, <strong>for</strong> the predicted static bending <strong>MOE</strong> <strong>and</strong> <strong>MOR</strong>, the coefficient of<br />

determination decreases by approximately half.


Practical grading into MGP grades<br />

Using best vibration prediction <strong>for</strong> logs<br />

For each resource, the best prediction equation from the resonance method was used to<br />

simulate the MGP grade recoveries.<br />

The removal of logs was based on a predicted threshold <strong>for</strong> <strong>MOE</strong> <strong>and</strong> <strong>MOR</strong>. The strategy was<br />

to target loss of no more than 5% of the potential MGP10 recovery through removal of logs.<br />

Plots <strong>for</strong> the most promising logs in each resource are compared in Figure 32 below which<br />

displays the percentage loss <strong>for</strong> each MGP grade <strong>and</strong> non-structural material. Because the<br />

MGP10 losses are st<strong>and</strong>ardised between the three resources, there is little difference in the<br />

MGP10 recovery.<br />

For Radiata E, 13.5% of the logs were removed using the 5% threshold <strong>for</strong> MGP10.<br />

Consequently, 32.1% of the non-structural material, 3.4% MGP10, <strong>and</strong> 0% <strong>for</strong> MGP12, were<br />

removed.<br />

For Radiata R, 4.6% of the logs were removed using the 5% threshold <strong>for</strong> MGP10.<br />

Consequently, 8.7% of the non-structural material, 3.3% MGP10, <strong>and</strong> 0% <strong>for</strong> MGP12 <strong>and</strong><br />

MGP15, were removed.<br />

For Caribbean, 5.9% of the logs were removed using the 5% threshold <strong>for</strong> MGP10.<br />

Consequently, 13.3% of the non-structural material, 4.7% MGP10, 0% <strong>for</strong> MGP12, <strong>and</strong> 1.2%<br />

MGP15 were removed.<br />

The MGP10 recovery of all three resources is similar as it was a key criterion <strong>for</strong> setting the<br />

threshold <strong>for</strong> rejection of logs.<br />

For each plot, the non-structural (NS) had the lowest recovery in the remaining pieces. The<br />

implications derived from these results will vary between processors <strong>and</strong> resource<br />

characteristics.<br />

62


100%<br />

90%<br />

80%<br />

70%<br />

60%<br />

50%<br />

40%<br />

30%<br />

20%<br />

10%<br />

0%<br />

NS MGP10 MGP12 MGP15 Logs lost<br />

63<br />

Radiata E<br />

Radiata R<br />

Caribbean<br />

Figure 32. Histogram of recovery based on r<strong>and</strong>om data by MGP grade when no more than 5% of the<br />

potential MGP10 is lost through removal of poor logs.<br />

The differences are in the percentage of logs removed <strong>and</strong> the recovery into NS in the<br />

remnant material. Radiata E had the best results with the largest reduction in NS after drying<br />

<strong>and</strong> planing. Radiata R had the worst results with the lowest reduction in NS. This may be<br />

linked to the good growing conditions <strong>for</strong> Radiata R, where bigger logs contain a large core of<br />

large knots, juvenile wood but a large proportion of the periphery contains good quality wood.<br />

The size of the log facilitates a high recovery of this part. With such good growth conditions<br />

the “mature” wood is proportionally higher than in the case of tree growing on poorer sites<br />

where the trees display a slower speed of growth when ageing, leading to smaller log where<br />

the good part is more difficult to recover. The size of the log is the key: the bigger the log, the<br />

more variety of wood there is in the log, so rejecting a whole log inevitably also rejects some<br />

good timber.<br />

Pre-sorting the cants (including the juvenile core) rather than whole logs, could be a better<br />

option <strong>for</strong> the resource, as the wing boards are likely to contain suitable material <strong>for</strong> structural<br />

grade products.<br />

Using best prediction <strong>for</strong> dry boards NDT <strong>technologies</strong><br />

We have restricted the recovery simulations on dry boards to two different combinations of<br />

equipment:<br />

• compression vibration plus WoodEye ® (WE Comp Vib);<br />

• compression vibration plus WoodEye ® plus Metriguard (WE Comp Vib Metriguard).<br />

PLS was used with the above combinations of grading <strong>technologies</strong> to develop predictions <strong>for</strong><br />

the biased test <strong>MOR</strong> from the grading parameters of each board tested. A separate PLS study


developed predictions of the biased test <strong>MOE</strong> <strong>for</strong> the same combinations of grading<br />

technology. Each of these predictions was then used to assign a grade to the whole board<br />

based on a threshold. The thresholds were evaluated by trying a number of different<br />

thresholds with the r<strong>and</strong>om position test data, until the pieces that fell into each grade gave<br />

characteristic values that exceeded the design characteristic value <strong>for</strong> all grades<br />

simultaneously. These thresholds could then be used to sort either the biased position test data<br />

or the r<strong>and</strong>om position test data into stress grades.<br />

Additionally, a composite grading was used in which each of the combinations used both the<br />

<strong>MOE</strong> prediction <strong>and</strong> the <strong>MOR</strong> prediction, with the requirement that the predictions <strong>for</strong> any<br />

board had to exceed both the <strong>MOE</strong> threshold <strong>and</strong> the <strong>MOR</strong> threshold <strong>for</strong> the grade to be<br />

admitted to the stress grade. The thresholds were set using the r<strong>and</strong>om position test data as<br />

indicated above.<br />

Figure 33 <strong>and</strong> Figure 34 display the correlations between the best grade parameter above <strong>and</strong><br />

<strong>MOE</strong> (Figure 32) a <strong>MOR</strong> (Figure 33) <strong>for</strong> the three resources when applied on biased test data.<br />

The main observation is the similarity between the correlations on these three different<br />

resources. This is expressed by a low difference on R 2 associated with a tiny difference on<br />

both slope <strong>and</strong> constant of the associated linear equation.<br />

MoE (MPa)<br />

25000<br />

20000<br />

15000<br />

10000<br />

5000<br />

0<br />

Radiata E<br />

Radiata R<br />

Caribbean<br />

Linear (Radiata E)<br />

Linear (Radiata R)<br />

Linear (Caribbean)<br />

64<br />

R² = 0.77<br />

R² = 0.77<br />

R² = 0.85<br />

0 5000 10000 15000 20000<br />

Grade parameter<br />

Figure 33. Correlation between <strong>MOE</strong> <strong>and</strong> grade parameter from the best combinations of WoodEye ® ,<br />

compression vibration <strong>and</strong> Metriguard predictors. Biased <strong>MOE</strong> <strong>and</strong> <strong>MOR</strong> threshold combination was<br />

applied.


MoR (MPa)<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

Radiata E<br />

Radiata R<br />

Caribbean<br />

Linear (Radiata E)<br />

Linear (Radiata R)<br />

Linear (Caribbean)<br />

65<br />

R² = 0.55<br />

R² = 0.49<br />

R² = 0.48<br />

0 2000 4000 6000 8000 10000 12000 14000 16000 18000 20000<br />

Grade parameter<br />

Figure 34. Correlation between <strong>MOR</strong> <strong>and</strong> grade parameter from the best combinations of WoodEye ® ,<br />

compression vibration <strong>and</strong> Metriguard predictors. Biased <strong>MOE</strong> <strong>and</strong> <strong>MOR</strong> threshold combination was<br />

applied.<br />

Figure 35, Figure 36, <strong>and</strong> Figure 37 display the coefficient of determination <strong>and</strong> the<br />

coefficient of variation <strong>for</strong> each resource. For a base-line comparison, two combinations <strong>and</strong><br />

Metriguard grading parameters applied to the biased position test data are presented.<br />

For the Radiata E resource, Figure 35, shows that the R 2 <strong>for</strong> all grading methods <strong>and</strong> <strong>MOE</strong> are<br />

close to the range of 80% to 90% which is expected to be the maximum attainable within the<br />

bounds of experimental error. For this resource, most <strong>MOE</strong> based grading methods were<br />

demonstrated to per<strong>for</strong>m effectively. There was little difference between the correlations of<br />

any of the combinations <strong>and</strong> the Metriguard grading.<br />

However, Figure 35 also shows that the R 2 <strong>for</strong> all of the combinations <strong>and</strong> <strong>MOR</strong> was better<br />

than that <strong>for</strong> the Metriguard grading. As this resource is primarily limited by its <strong>MOE</strong><br />

properties, there may be a large number of <strong>MOE</strong> measurement devices (including the<br />

Metriguard) that could return grading as effective as the best of the combination systems. It<br />

also shows that there is little difference in the COV of bending strength of each of the<br />

commercial grades <strong>and</strong> hence all of the plotted grading methods return comparable reliability<br />

of <strong>graded</strong> product.


Coefficient of determinationn<br />

Strength Coefficient of Variation <strong>for</strong> biased test<br />

1<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

60%<br />

50%<br />

40%<br />

30%<br />

20%<br />

10%<br />

0%<br />

0<br />

<strong>MOE</strong> <strong>MOR</strong><br />

Correlation coefficient<br />

66<br />

<strong>MOR</strong> WE Comp Vib B<br />

<strong>MOE</strong> WE Comp Vib B<br />

Composite 2 B<br />

<strong>MOR</strong> WE Comp Vib Metriguard B<br />

<strong>MOE</strong> WE Comp Vib Metriguard B<br />

Composite 3 B<br />

Min Metriguard B<br />

NS MGP10 MGP12<br />

Grade<br />

Figure 35. Radiata E resource, coefficient of determination <strong>and</strong> coefficient of variation <strong>for</strong> strength<br />

when grading with the best combinations of WoodEye ® , compression vibration <strong>and</strong> Metriguard.<br />

Biased <strong>MOE</strong>, <strong>MOR</strong> or combination (composite) of both thresholds was applied. Note: WE Comp Vib<br />

= 2 <strong>and</strong> WE Comp Vib Metriguard = 3.<br />

For Radiata R resource, Figure 36 below shows there is a minor improvement in R 2 with<br />

<strong>MOE</strong> using the <strong>MOE</strong> predictions or the Composite <strong>for</strong> both combinations. There is a similar<br />

improvement in R 2 with <strong>MOR</strong> using the <strong>MOR</strong> predictions or the Composite <strong>for</strong> both<br />

combinations. There was little variation in the COV of <strong>MOR</strong> <strong>for</strong> <strong>graded</strong> material between the<br />

grading methods with comparable reliability of <strong>graded</strong> product <strong>for</strong> all of these grading<br />

methods.


Coefficient of determinationn<br />

Strength Coefficient of Variation <strong>for</strong> biased test<br />

1<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

60%<br />

50%<br />

40%<br />

30%<br />

20%<br />

10%<br />

0%<br />

0<br />

<strong>MOE</strong> <strong>MOR</strong><br />

Correlation coefficient<br />

67<br />

<strong>MOR</strong> WE Comp Vib B<br />

<strong>MOE</strong> WE Comp Vib B<br />

Composite 2 B<br />

<strong>MOR</strong> WE Comp Vib Metriguard B<br />

<strong>MOE</strong> WE Comp Vib Metriguard B<br />

Composite 3 B<br />

Min Metriguard B<br />

NS MGP10 MGP12<br />

Grade<br />

Figure 36. Radiata R resource, coefficient of determination <strong>and</strong> coefficient of variation <strong>for</strong> strength<br />

when grading with the best combinations of WoodEye ® , compression vibration <strong>and</strong> Metriguard.<br />

Biased <strong>MOE</strong>, <strong>MOR</strong> or combination (composite) of both thresholds was applied. Note: WE Comp Vib<br />

= 2 <strong>and</strong> WE Comp Vib Metriguard = 3.<br />

For Caribbean resource the use of a combination of grading equipment or predictors<br />

significantly increased the R 2 of both <strong>MOE</strong> <strong>and</strong> <strong>MOR</strong>. The R 2 with <strong>MOE</strong> was increased most<br />

using the <strong>MOE</strong> prediction <strong>and</strong> the R 2 with <strong>MOR</strong> was increased most using the <strong>MOR</strong><br />

prediction. These <strong>and</strong> the composite grading returned higher R 2 than the base-line Metriguard<br />

grading <strong>for</strong> both <strong>MOE</strong> <strong>and</strong> <strong>MOR</strong> as shown in Figure 37.


The COV of the <strong>MOR</strong> <strong>for</strong> each of the MGP12 <strong>and</strong> MGP15 grades, also shown in Figure 37,<br />

showed a significant reduction below its Metriguard base-line value with the application of<br />

new technology. This resource offered the most potential <strong>for</strong> improvement of grading using a<br />

combination of grading <strong>technologies</strong>.<br />

Coefficient of determinationn<br />

Strength Coefficient of Variation <strong>for</strong> biased test<br />

1<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

60%<br />

50%<br />

40%<br />

30%<br />

20%<br />

10%<br />

0<br />

<strong>MOE</strong> <strong>MOR</strong><br />

Correlation coefficient<br />

68<br />

<strong>MOR</strong> WE Comp Vib B<br />

<strong>MOE</strong> WE Comp Vib B<br />

Composite 2 B<br />

<strong>MOR</strong> WE Comp Vib Metriguard B<br />

<strong>MOE</strong> WE Comp Vib Metriguard B<br />

Composite 3 B<br />

Min Metriguard B<br />

0%<br />

NS MGP10 MGP12 MGP15<br />

Grade<br />

Figure 37. Caribbean resource, coefficient of determination <strong>and</strong> coefficient of variation <strong>for</strong> strength<br />

when grading with the best combinations of WoodEye ® , compression vibration <strong>and</strong> Metriguard.<br />

Biased <strong>MOE</strong>, <strong>MOR</strong> or combination (composite) of both thresholds was applied. Note: WE Comp Vib<br />

= 2 <strong>and</strong> WE Comp Vib Metriguard = 3.


Recoveries based on r<strong>and</strong>om test<br />

70%<br />

60%<br />

50%<br />

40%<br />

30%<br />

20%<br />

10%<br />

0%<br />

NS MGP10 MGP12 MGP15<br />

Grade<br />

Figure 40 display the recoveries from the application of the same grade thresholds illustrated above to<br />

the full suite of boards represented by the r<strong>and</strong>om data.<br />

69<br />

<strong>MOR</strong> WE Comp Vib R<br />

<strong>MOE</strong> WE Comp Vib R<br />

Composite 2 R<br />

<strong>MOR</strong> WE Comp Vib Metriguard R<br />

<strong>MOE</strong> WE Comp Vib Metriguard R<br />

Composite 3 R<br />

Min Metriguard R<br />

The important thing to focus on in these graphs is the recovery to non-structural (NS). We<br />

want to minimise this as this is timber that loses money <strong>for</strong> the mill. The tendencies observed<br />

on the quality regression indicators when applying the best predictions on biased test data are<br />

confirmed on the recoveries while applying on r<strong>and</strong>om test data, namely:<br />

• Radiata E resource: there is no improvement observed when using the best<br />

combinations of WoodEye ® , compression vibration <strong>and</strong> Metriguard compared to the<br />

use of Metriguard alone.<br />

• Radiata R resource: the recovery increase observed on MGP10 grade is somewhat<br />

offset by an increase of non structural products <strong>and</strong> a recovery decrease in MGP 12<br />

grade.<br />

• Caribbean resource: a significant recovery increase on MGP10 <strong>and</strong> MGP15 grades<br />

associated with an important decrease of non structural products. Despite an absolute<br />

decrease of 15% on MGP12 recovery, the 10% absolute decrease of non structural<br />

products associated with a 20% absolute increase on MGP10 recovery represents<br />

certainly a key advantage <strong>for</strong> this resource.


70%<br />

60%<br />

st<br />

te<br />

m<br />

o 50%<br />

d<br />

n<br />

ra<br />

n 40%<br />

o<br />

d<br />

b<br />

a<br />

se<br />

30%<br />

s<br />

v<br />

e<br />

rie<br />

20%<br />

e<br />

c<br />

o<br />

R<br />

10%<br />

0%<br />

NS MGP10 MGP12 MGP15<br />

Grade<br />

Figure 38. Radiata E recoveries from calibrations developed with biased test data <strong>for</strong> each resource<br />

<strong>and</strong> <strong>for</strong> different grading parameters when using static bending r<strong>and</strong>om data. Biased <strong>MOE</strong>, <strong>MOR</strong> or<br />

combination (composite) of both thresholds was applied. Note: WE Comp Vib = 2 <strong>and</strong> WE Comp Vib<br />

Metriguard = 3.<br />

70<br />

<strong>MOR</strong> WE Comp Vib R<br />

<strong>MOE</strong> WE Comp Vib R<br />

Composite 2 R<br />

<strong>MOR</strong> WE Comp Vib Metriguard R<br />

Radiata E <strong>MOE</strong> WE Comp Vib Metriguard R<br />

Composite 3 R<br />

Min Metriguard R


Recoveries based on r<strong>and</strong>om test<br />

70%<br />

60%<br />

50%<br />

40%<br />

30%<br />

20%<br />

10%<br />

0%<br />

NS MGP10 MGP12 MGP15<br />

Grade<br />

Figure 39. Radiata R recoveries from calibrations developed with biased test data <strong>for</strong> each resource<br />

<strong>and</strong> <strong>for</strong> different grading parameters when using static bending r<strong>and</strong>om data. Biased <strong>MOE</strong>, <strong>MOR</strong> or<br />

combination (composite) of both thresholds was applied. Note: WE Comp Vib = 2 <strong>and</strong> WE Comp Vib<br />

Metriguard = 3.<br />

Recoveries based on r<strong>and</strong>om test<br />

70%<br />

60%<br />

50%<br />

40%<br />

30%<br />

20%<br />

10%<br />

0%<br />

71<br />

<strong>MOR</strong> WE Comp Vib R<br />

<strong>MOE</strong> WE Comp Vib R<br />

Composite 2 R<br />

<strong>MOR</strong> WE Comp Vib Metriguard R<br />

<strong>MOE</strong> WE Comp Vib Metriguard R<br />

Composite 3 R<br />

Min Metriguard R<br />

<strong>MOR</strong> WE Comp Vib R<br />

<strong>MOE</strong> WE Comp Vib R<br />

Composite 2 R<br />

<strong>MOR</strong> WE Comp Vib Metriguard R<br />

<strong>MOE</strong> WE Comp Vib Metriguard R<br />

Composite 3 R<br />

Min Metriguard R<br />

NS MGP10 MGP12 MGP15<br />

Grade<br />

Figure 40. Caribbean recoveries from calibrations developed with biased test data <strong>for</strong> each resource<br />

<strong>and</strong> <strong>for</strong> different grading parameters when using static bending r<strong>and</strong>om data. Biased <strong>MOE</strong>, <strong>MOR</strong> or<br />

combination (composite) of both thresholds was applied. Note: WE Comp Vib = 2 <strong>and</strong> WE Comp Vib<br />

Metriguard = 3.


Discussion <strong>and</strong> conclusions<br />

The capability of a grading system to predict accurately <strong>and</strong> reliably the reference parameters<br />

of a product, namely the board’s actual stiffness <strong>and</strong> strength, depends mainly on how well<br />

the prediction agrees with:<br />

• the reference parameters (indicated by R 2 , the coefficient of determination);<br />

• the measurement error of the predictor parameter(s) (indicated by COV, the<br />

coefficient of variation); <strong>and</strong><br />

• the capability of the system to accommodate resource variability (indicated by its<br />

prediction robustness <strong>and</strong> the level of recovery compared to the actual grading based<br />

on destructive tests).<br />

If the regression analysis is based on measurements made in the same conditions <strong>and</strong> the same<br />

apparatus that is used in grading machines, the effect of the measurement error <strong>and</strong> COV is<br />

already included in the R 2 value directly. This was the case <strong>for</strong> the in-line equipment tested.<br />

For the resonance methods made in laboratory conditions with the Bing ® equipment, the effect<br />

of measurement error should be considered separately, when evaluating its effectiveness.<br />

However resonance methods have proven to be very robust <strong>and</strong> accurate due to their<br />

metrological 1 simplicity <strong>and</strong> fundamental principles. There is a large quantity of publications<br />

based on theoretical mechanics work <strong>and</strong> experimental results which demonstrate the<br />

capability to predict stiffness <strong>and</strong> strength parameters of wood <strong>and</strong> other materials. In this<br />

study the comparison of results obtained between in-line equipment (Thumper <strong>and</strong><br />

Metriguard) <strong>and</strong> a laboratory system (Bing ® ) demonstrated the high degree of closeness<br />

between in-line <strong>and</strong> laboratory measurements on one h<strong>and</strong>, <strong>and</strong> the quality of the prediction of<br />

the reference static bending <strong>MOE</strong> <strong>and</strong> <strong>MOR</strong> on the other h<strong>and</strong>. Theoretical, flexion should<br />

provide the best results due to the parallels to static bending, that is the configuration implied<br />

flexure stress. However, due to metrological constraints, low frequencies <strong>and</strong> signal duration,<br />

plus boundary conditions, compression <strong>and</strong> flexion vibration provide the same level of<br />

prediction.<br />

The sample size <strong>for</strong> each resource (approximately five hundred boards per resource) was large<br />

enough to make the results reliable. The measurements of the various parameters of this study<br />

were per<strong>for</strong>med at different times <strong>and</strong> sometimes under dissimilar conditions. Despite the<br />

ef<strong>for</strong>t to keep these conditions as homogenous as possible some variations may persist. As a<br />

consequence, the results presented here should not be used as definitive when the difference<br />

between the R 2 is typically less than 0.02.<br />

As was found in the European Combigrade projects’ (Hanhijärvi et al. 2008 <strong>and</strong> 2005) results<br />

<strong>for</strong> dry boards, this study demonstrates that the best single parameter predictors of <strong>MOE</strong> <strong>and</strong><br />

<strong>MOR</strong> are the stiffness related measurements derived by either direct mechanical test<br />

(Metriguard) or vibration method. These two approaches offer a similar level of prediction<br />

which the R 2 ranges from 0.66 to 0.85 <strong>for</strong> <strong>MOE</strong> <strong>and</strong> 0.35 to 0.53 <strong>for</strong> <strong>MOR</strong>, depending on the<br />

resource. Importantly because the Metriguard provides both average <strong>and</strong> local board<br />

measurements it shows better per<strong>for</strong>mance on strength limited or defect affected resources.<br />

On clearly stiffness-limited resource, vibration seems to provide a slightly better prediction<br />

than Metriguard. On this type of resource combining stiffness parameters with local<br />

measurements, knot or local stiffness, does not improve the prediction significantly.<br />

1 relating to metrology. Metrology is the scientific study of measurement.<br />

72


Reference static bending parameters prediction <strong>for</strong> strength limited, or defect affected<br />

resources, can be notably improved by combining stiffness parameters with local in<strong>for</strong>mation<br />

<strong>and</strong> an R 2 improvement magnitude up to 0.1 <strong>for</strong> <strong>MOE</strong> <strong>and</strong> 0.2 <strong>for</strong> <strong>MOR</strong> can be expected.<br />

The choice of the statistical approach to achieve these improvements is determinant. The<br />

application of classical MLR on several predictors proved to be effective, in particular <strong>for</strong><br />

<strong>MOR</strong> predictions. However the usage of a modern statistics approach (chemometrics tools)<br />

such as PLS proves to be more efficient <strong>for</strong> <strong>improving</strong> the level of prediction. This is due to<br />

the problem of multi-collinearity in MLR which impaired the efficient exploitation of the<br />

in<strong>for</strong>mation contained in all the predictors available <strong>for</strong> a given combination of grading<br />

systems. As the development of robust calibration is based on training samples (samples<br />

specifically selected to develop the prediction equation) <strong>and</strong> application of complex<br />

mathematics, the robustness <strong>and</strong> accuracy of the prediction requires a sound knowledge <strong>and</strong> a<br />

disciplined approach.<br />

Obviously statistical analysis alone is only a criterion, even if decisive, when considering the<br />

fitness of a grading system to a certain application. The price of the system, its maintenance<br />

cost, its production line adaptation, versatility, reliability, simplicity, etc. are all important<br />

factors.<br />

Moreover, the reference method accuracy is not perfect. The impact of the experimental error<br />

on the reference’s actual parameters leads to a maximum attainable level of prediction within<br />

the bounds of error. When the quality of the prediction comes closer to this threshold, any<br />

improvement requires much more ef<strong>for</strong>t <strong>and</strong> consequently investment. The balance between<br />

grade selection accuracy <strong>and</strong> investment depend on each processor’s strategy.<br />

The consequence of the application of combined grading systems on grade recoveries has to<br />

be analysed carefully since it is clearly resource dependent. The trend observed shows that<br />

there is no advantage to using combined grading systems <strong>for</strong> a stiffness-limited resource. The<br />

advantage <strong>for</strong> a strength limited resource has to be weighed with other processing impacts.<br />

The Caribbean resource offers the most promising improvement potential using a<br />

combination of new <strong>technologies</strong>.<br />

The use of resonance method <strong>for</strong> <strong>MOE</strong> <strong>and</strong> <strong>MOR</strong> prediction on green boards provides slightly<br />

less successful results when compared to dry boards. Nevertheless, the R 2 decrease is only 0.1<br />

to 0.15 depending on the resource. This result is a real opportunity <strong>for</strong> processors to try to<br />

apply an early detection of the weakest boards in order to relieve the cost of non-value added<br />

further processing. Moreover the potential use of optical scanners or equivalent equipment<br />

allowing the detection of grain deviation <strong>and</strong> knot characteristics on green boards promises to<br />

increase quite significantly the <strong>MOR</strong> prediction. It represents an avenue to improve the<br />

efficiency of the mill.<br />

Log segregation through vibration method (compression) is also resource dependent. From<br />

the results of this study, it is clear that the stiffness-limited resource should have the highest<br />

gain in recovery, through removal of the lowest stiffness logs. There<strong>for</strong>e, significant cost<br />

savings could be achieved by processing these into non-structural products, with minimal<br />

impact on final <strong>graded</strong> product recovery. The value gain from log stiffness sorting is less <strong>for</strong><br />

Caribbean pine <strong>and</strong> requires further <strong>assessment</strong>. For strength limited resources, the<br />

consequences on recovery are insignificant.<br />

73


Recommendations<br />

All the results obtained from this study should be validated with full in-mill simulations. This<br />

could be achieved using calibrations developed during this project <strong>and</strong> applied to independent<br />

samples in sufficient numbers of replicates covering the range of variability of material<br />

h<strong>and</strong>led by each processor.<br />

There is room <strong>for</strong> improvement on each system used, <strong>for</strong> example the use of an optical system<br />

could be tuned <strong>for</strong> the resource’s grade limiting features. The WoodEye ® system used in this<br />

study was tuned <strong>for</strong> appearance grading of hoop pine <strong>and</strong> it was outside the scope of the<br />

project to re-tune the equipment <strong>for</strong> structural grading purposes. Moreover, the raw data<br />

variables extracted from WoodEye ® weren’t refined to provide the most appropriate<br />

characterization of the material tested. Further discussion with qualified optical scanning<br />

system engineers should provide optimized extraction <strong>for</strong> structural grading purposes. This<br />

will involve the development of clear definitions of the optical characteristics that impact<br />

<strong>MOE</strong> <strong>and</strong> <strong>MOR</strong> <strong>for</strong> each resource.<br />

Vibration method can be improved through:<br />

• installation of devices at strategic locations in the mill (i.e. cants);<br />

• refining data extracted, e.g. using current dataset, further analysis to improve<br />

algorithm extraction;<br />

• using flexure with different boundary conditions, e.g. testing local span within boards.<br />

At green board level, an optical scanning system coupled with a vibration system, has proved<br />

to be nearly as efficient as the same combination on dry boards. Depending on the processor’s<br />

strategy, further investigation may provide an avenue <strong>for</strong> efficient green mill grading<br />

<strong>technologies</strong>.<br />

A larger sample of logs should be tested to verify the potential <strong>for</strong> effective log segregation.<br />

An error analysis on the reference method <strong>for</strong> <strong>MOE</strong>/<strong>MOR</strong> measurement should be per<strong>for</strong>med<br />

in order to assess the maximal level of prediction achievable.<br />

Locating vibration devices at strategic locations in the mill may provide a cost-effective<br />

alternative to investment in higher cost capital items installed near to the end of the process<br />

line.<br />

A further project could be designed to continue mining the existing dataset. For example,<br />

testing the consequence of grading at stages through the flow of the boards.<br />

74


References<br />

Baradit, E., Aedo R. <strong>and</strong> Correa J. (2006). Knots detection in wood using microwaves, Wood<br />

Science <strong>and</strong> Technology 40 (pp118-123).<br />

Bendtsen B.A. (1978) Properties of wood from improved <strong>and</strong> intensively managed trees.<br />

Bootle K.R., (2005). WOOD IN AUSTRALIA. TYPES, PROPERTIES AND USES. 2nd Ed.<br />

McGraw-Hill Book Company, Sydney, Australia.<br />

Brancheriau, L. (2002). Expertise mécanique des sciages par analyses des vibrations dans le<br />

domaine acoustique, Université de la méditerranée - Aix Marseille II, Université de la<br />

méditerranée - Aix Marseille II, Marseille (p 278).<br />

Brancheriau, L. (2007). WISIS - Wood In Situ Inspection- Instructions <strong>for</strong> use Version 1.0.<br />

CIRAD.<br />

Brancheriau, L. <strong>and</strong> Bailleres, H. (2002). Natural vibration analysis of clear wooden beams: a<br />

theoretical review. Wood Science <strong>and</strong> Technology 36 (pp 346-365). Springer-Verlag.<br />

Brancheriau, L. <strong>and</strong> Bailleres, H. (2003). Use of Partial Least Squares Method with Acoustic<br />

Vibration Spectra as a New grading Technique <strong>for</strong> Structural Timber. Holz<strong>for</strong>schung Vol. 57,<br />

No. 6 (pp 644-652).<br />

Bucur, V. (2003). NONDESTRUCTIVE CHARACTERIZATION AND IMAGING OF<br />

WOOD. Springer-Verlag, Berlin, Germany.<br />

Duff, G. (2005). Technology <strong>for</strong> delivering high quality <strong>graded</strong> softwood product- practical<br />

applications. Gottstein Fellowship report.<br />

Grabianowski, M., B. Manley, <strong>and</strong> J.C.F. Walker. 2006. Acoustic measurements on st<strong>and</strong>ing<br />

trees, logs <strong>and</strong> green lumber. Wood Sci. Technol. 40(3):205-216.<br />

Halabe, U.B., Bidigalu G.M., GangaRao H.V.S. <strong>and</strong> Ross R.J. (1995). Nondestructive<br />

Evaluation of Green Wood Using Stress Wave <strong>and</strong> Transverse Vibration Techniques. Vol. 55,<br />

No. 9.<br />

Hanhijärvi, A. <strong>and</strong> Ranta-Maunus, A (2008). Development of strength grading of timber<br />

using combined measurement techniques. Report of the Combigrade project – phase 2. VTT<br />

Publications 686.<br />

Hanhijärvi, A., Ranta-Maunus, A <strong>and</strong> Turk, G. (2005). Potential of strength grading of timber<br />

with combined measurement techniques. Report of the Combigrade project – phase 1. VTT<br />

Publications 568.<br />

Harris J.M. (1991). PROPERTIES AND USES OF NEW ZEALAND RADIATA PINE.<br />

Volume one - wood properties, Forest Research Institute, Rotorua, New Zeal<strong>and</strong>.<br />

Hopewell, G. (Ed.) 2006. CONSTRUCTION TIMBERS IN QUEENSLAND: PROPERTIES<br />

AND SPECIFICATIONS FOR SATISFACTORY PERFORMANCE OF CONSTRUCTION<br />

75


TIMBERS IN QUEENSLAND, CLASS 1 AND CLASS 10 BUILDINGS (BOOKS 1 AND<br />

2). Department of Primary Industries <strong>and</strong> Fisheries, Brisbane.<br />

Ilic, J., Northway, R. <strong>and</strong> Pongracic, S. (2003). Juvenile Wood Characteristics, Effects <strong>and</strong><br />

Identification. Literature Review. FWPRDC Project No. PN02.1907.<br />

Kininmonth, J.A., Whitehouse, L.J. (1991). Properties <strong>and</strong> Uses of New Zeal<strong>and</strong> Radiata<br />

Pine, Vol. 1. Forest Research Institute, Rotorua, New Zeal<strong>and</strong>,<br />

Krautkrämer, J. K.H. (1986). Werkstoffprüfung mit Ultraschall, Springer Verlag, Berlin.<br />

Krzosek, S. (2003). Machine stress-grading of timber: current status <strong>and</strong> prospects of<br />

development, Festigkeitssortierung des Bauholzes mit Maschinen - Ist-St<strong>and</strong> und<br />

Entwicklungsperspektiven.<br />

Lamb A.F.A., (1973). Pinus caribaea FAST GROWING TREES OF THE LOWLAND<br />

TROPICS Volume 1. N°6. University of Ox<strong>for</strong>d, Engl<strong>and</strong> (254 pp).<br />

Leicester, R.H. (2004). Grading <strong>for</strong> structural timber. Progress in Structural engineering <strong>and</strong><br />

Materials. Vol. 6 (pp 69-78).<br />

Leicester, R.H., Breitinger, H.O. <strong>and</strong> Fordham, H.F. (1996). Equivalence of in-grade testing<br />

st<strong>and</strong>ards. CIB/W18A- Timber Structures, Paper 29-6-2, Bordeaux, France.<br />

Ross, R.J. <strong>and</strong> Pellerin, R.F. (1994). Non-destructive testing <strong>for</strong> assessing wood members in<br />

structures- a review. (40 pp). Rev. Ed. Forest Products Laboratory USDA Forest Service,<br />

Madison USA.<br />

Ross, R.J., Brashaw B.K. <strong>and</strong> Wang X.P. (2006). Structural condition <strong>assessment</strong> of inservice<br />

wood. Forest Products Journal 56, 4.<br />

Ross, R.J., Zerbe, J.I., Wang, X.P., Green D.W. <strong>and</strong> Pellerin R.F. (2005). Stress wave<br />

nondestructive evaluation of Douglas-fir peeler cores. Forest Products Journal 55 (pp 90-94).<br />

S<strong>and</strong>oz, J.L. (1989). Grading of construction timber by ultrasound. Wood Science <strong>and</strong><br />

Technology 23 (pp 95-108).<br />

Schimleck, L.R., Evans, R., Ilic, J. <strong>and</strong> Matheson A.C. (2002). Estimation of wood stiffness of<br />

increment cores by near-infrared spectroscopy. Canadian Journal of Forest Research 32 (pp<br />

129-135).<br />

Siemon, G. (1981). Effect of CCA-preservative treatment on bending strength of small clear<br />

specimens of high-temperature dried <strong>and</strong> air-dried Caribbean pine. Technical Paper No. 27. (4<br />

pp). Department of Forestry, Queensl<strong>and</strong>.<br />

Smith, R.G.B., Palmer, G. Davies, M. And Muneri, A (2003). A method enabling the<br />

reconstruction of internal features of logs from sawn lumber: the log end template. Forest<br />

Products Journal Volume 53 No. 11/12 (pp 95-98).<br />

Steiger, R. (1996). Mechanische Eigenschaften von Schweizer Fichten-Bauholz bei Biege-,<br />

Zug-, Druck- und kombinierter M/N-Belastung Sortierung von Rund- und Schnittholz mittels<br />

76


Ultraschall, VII, 168S. Institut für Baustatik und Konstruktion (IBK), Eidgenössische<br />

Technische Hochschule Zürich (ETH), Zürich.<br />

Tsehaye, A.., Buchanan, A.H. <strong>and</strong> Walker, J.C.F. (2000). Sorting of logs using acoustics.<br />

Wood Science <strong>and</strong> Technology 34 (pp 337-344).<br />

Tsuchikawa, S. (2007). A Review of Recent Near Infrared Research <strong>for</strong> Wood <strong>and</strong> Paper.<br />

Applied Spectroscopy Reviews 42 (1) (pp 43–71).<br />

Wang, X.P., Ross, R.J., Green, D.W., Brasha, B., Englund, K. <strong>and</strong> Wolcott, M. (2003). Stress<br />

wave sorting of red maple logs <strong>for</strong> structural quality. Wood Science <strong>and</strong> Technology 37 (pp<br />

531-537).<br />

Yeh, T.F., Chang, H.M. <strong>and</strong> Kadla, J.F. (2004). Rapid prediction of solid wood lignin content<br />

using transmittance near-infrared spectroscopy. Journal of Agricultural <strong>and</strong> Food Chemistry<br />

52 (pp 1435-1439).<br />

Zobel, B.J. <strong>and</strong> Buijtenen J.P.V (1989). WOOD VARIATION: ITS CAUSES AND<br />

CONTROL. Springer-Verlag, Berlin, Germany.<br />

Internet references<br />

Web 1. http://www.weeds.org.au/cgi-bin/weedident.cgi?tpl=plant.tpl&ibra=all&card=E41#<br />

Visited in April 2007<br />

Web 2. http://www.statsoft.com/textbook<br />

Visited 31.07.09<br />

Web 3. http://www2.dpi.qld.gov.au/hardwoodsqld/13322.html<br />

Visited in August 2008.<br />

Web 4. http://www.metriguard.com/<br />

Visited in August 2008<br />

Web 5. http://www.newnes-mcgehee.com/splash.htm<br />

Visited in August 2008<br />

St<strong>and</strong>ards<br />

AS/NZS 1748:2003 - Timber-Stress-<strong>graded</strong>-product requirements <strong>for</strong> mechanically stress<strong>graded</strong><br />

timber.<br />

EN 408 (1995). European St<strong>and</strong>ard 408, Timber structures- Structural timber <strong>and</strong> glued<br />

laminated timber- Determination of some physical <strong>and</strong> mechanical properties. Brussels.<br />

EN 14081-1 (2003). European St<strong>and</strong>ard E. Timber structures- Strength <strong>graded</strong> structural<br />

timber with rectangular cross section- Part 1: General requirements. Brussels.<br />

AS 1720.1 (1997) Australian St<strong>and</strong>ard. Timber structures, Part 1: Design methods.<br />

77


AS/NZS 4063 (1992). Australian/New Zeal<strong>and</strong> St<strong>and</strong>ard 4063. Timber−Stress-<strong>graded</strong>−Ingrade−strength<br />

<strong>and</strong> stiffness evaluation.<br />

Personal communication<br />

Boughton, G. (2009). pers. comm. Director, TimberED, Bunbury Western Australia.<br />

Nikles, G. Dr. (2009). Research Associate, Queensl<strong>and</strong> Primary Industries <strong>and</strong> Fisheries.<br />

78


Acknowledgements<br />

The successful conduct of this project was made possible through the enthusiastic support<br />

from management <strong>and</strong> staff of the following organisations:<br />

• Carter Holt Harvey, Myrtle<strong>for</strong>d Victoria<br />

• Carter Holt Harvey, Caboolture Queensl<strong>and</strong> (<strong>for</strong>merly Weyerhaueser)<br />

• Wespine, Bunbury Western Australia<br />

• Hyne Timber, Tumbarumba New South Wales <strong>and</strong> Tuan Queensl<strong>and</strong><br />

• Queensl<strong>and</strong> Primary Industries <strong>and</strong> Fisheries, Brisbane Queensl<strong>and</strong><br />

• TimberED, Perth Western Australia<br />

• CIRAD Xylometry, Montpellier France<br />

79


Researcher’s Disclaimer<br />

© The State of Queensl<strong>and</strong>, Primary Industries <strong>and</strong> Fisheries, 2009.<br />

Except as permitted by the Copyright Act 1968, no part of the work may in any <strong>for</strong>m or by<br />

any electronic, mechanical, photocopying, recording, or any other means be reproduced,<br />

stored in a retrieval system or be broadcast or transmitted without the prior written permission<br />

of DPI&F. The in<strong>for</strong>mation contained herein is subject to change without notice. The<br />

copyright owner shall not be liable <strong>for</strong> technical or other errors or omissions contained herein.<br />

The reader/user accepts all risks <strong>and</strong> responsibility <strong>for</strong> losses, damages, costs <strong>and</strong> other<br />

consequences resulting directly or indirectly from using this in<strong>for</strong>mation.<br />

80


Appendix 1. Resonance method extracted signal descriptors<br />

SCGravity<br />

Spectral centre of gravity divided by the fundamental frequency (f1) in %:<br />

SCGf1 = SCG ./ f1 .* 100;<br />

SB<strong>and</strong>width<br />

with f(N) = Fmax range<br />

Spectral extent divided the fundamental frequency (f1) in %:<br />

Sb<strong>and</strong>f1 = Sb<strong>and</strong> ./ f1 .* 100;<br />

SSlope<br />

Spectral slope divided the fundamental frequency (f1) in %:<br />

Sslopef1 = Sslope ./ f1 .* 100;<br />

IF<br />

Inharmonicity factor:<br />

ME<br />

Average of the first three dynamic <strong>MOE</strong><br />

<strong>MOE</strong>_n<br />

Dynamic <strong>MOE</strong> associated with fn<br />

FacQ_n<br />

Quality factor (inverse internal damping or friction) associated with fn:<br />

81


Pow_n<br />

such as<br />

Mean power in frequency sub-b<strong>and</strong> (between 0 <strong>and</strong> f1; f1 <strong>and</strong> f2...):<br />

MSNrjRatioF_n<br />

Sub-b<strong>and</strong> energy ratio (between the resonance frequencies) :<br />

PowF_n<br />

= Sub-b<strong>and</strong> energy / Total range energy<br />

Mean power of the sub-b<strong>and</strong> (resonance shoulder defined by a pass b<strong>and</strong> of -20dB) :<br />

MSRatioF_n<br />

Sub-b<strong>and</strong> energy ratio (resonance shoulder defined by a pass b<strong>and</strong> of -20dB)<br />

= sub-b<strong>and</strong> energy / total range energy<br />

82


CTR<br />

Appendix 2. WoodEye profile descriptions.<br />

CTR_AC_M<br />

CTR_AC_5<br />

CTR_BD_5<br />

CTR_trgl_BD_5<br />

CTR_trgl_AC_M<br />

CTR_trgl_AC_3<br />

CTR_trgl_AC_5<br />

______________<br />

KAR<br />

KAR_ext_M<br />

KAR_ext_99<br />

KAR_ext_98<br />

KAR_int_M<br />

KAR_int_99<br />

KAR_trgl_M<br />

KAR_trgl_99<br />

Material inertia ratio = calculated inertia / beam inertia (b.h³/12)<br />

The inertia is calculated from the sum of inertias according to Parallel Axis Theorem<br />

(also known as Huygens-Steiner theorem)<br />

Profile is calculated between the external loading points<br />

Profile is weighted by the bending moment (unit maximum moment)<br />

Profile is equal to the product of the profiles from the 4 beam sides.<br />

Sides A&C, rectangular window, mean value<br />

Sides A&C, rectangular window, 5% fractile (equivalent to a minimum)<br />

Ratio defect surface/total surface; B&D edges, rectangular window, 5% fractile<br />

B&D edges, triangular window, 5% fractile<br />

Sides A&C, triangular window, mean value<br />

Sides A&C, triangular window, 3% fractile<br />

Sides A&C, triangular window, 5% fractile<br />

____________________________________________________________<br />

Perimeter ratio = calculated perimeter / beam perimeter (2.b.h)<br />

The perimeter is associated to defect height (sum of the heights)<br />

Profile is calculated between the external loading points<br />

Profile is equal to the sum of the profiles from the 4 beam sides.<br />

A&B&C&D, 1/3 external height <strong>for</strong> A&C), rectangular window, mean value<br />

99% fractile (equivalent to a maximum)<br />

98% fractile<br />

A&B&C&D, 1/3 internal height <strong>for</strong> A&C), rectangular window, mean value<br />

99% fractile<br />

A&B&C&D, triangular window, mean value<br />

99% fractile<br />

Notes: A <strong>and</strong> C are the faces <strong>and</strong> B <strong>and</strong> D are the board edges.<br />

83


Appendix 3. Biased <strong>and</strong> r<strong>and</strong>om testing protocol.<br />

Radiata E resource<br />

• Identify the limiting feature (that is, the largest strength limiting defect) to provide the<br />

biased sample location.<br />

• Mark the 1.8 m length so that this point lies within the middle third of the specimen.<br />

• Using a r<strong>and</strong>om number calculator, provide the distance from the board end <strong>for</strong><br />

r<strong>and</strong>om sample datum point.<br />

• If this position is situated so that a second 1.8 m long test piece can be created without<br />

overlapping the biased specimen, it is marked ‘true r<strong>and</strong>om’.<br />

• Otherwise a sample is cut directly next to the biased one <strong>and</strong> marked ‘false r<strong>and</strong>om’.<br />

• When the biased <strong>and</strong> the r<strong>and</strong>om samples are situated so that both conditions were<br />

satisfied, the specimen is marked with ‘biased <strong>and</strong> r<strong>and</strong>om’.<br />

• The specimens called true <strong>and</strong> false r<strong>and</strong>om are both considered as r<strong>and</strong>om samples.<br />

• The results obtained from the ‘biased <strong>and</strong> r<strong>and</strong>om’ samples are taken into account in<br />

r<strong>and</strong>om as well as biased analyses.<br />

This sampling methodology, whereby the biased sampling had priority <strong>and</strong> a r<strong>and</strong>om sample<br />

was only available if located outside the biased sample region, provided a r<strong>and</strong>om sample set<br />

of approximately 40% of the whole sub-sample.<br />

Radiata R <strong>and</strong> Caribbean resources<br />

• Using a r<strong>and</strong>om number calculator, provide the distance from the numbered board end<br />

<strong>for</strong> the r<strong>and</strong>om sample datum point.<br />

• A 1800 mm test specimen template with the centre third demarcated is placed at this<br />

datum to locate r<strong>and</strong>om specimen location.<br />

• Check to see if the maximum strength reducing defect is located in the centre third of<br />

the possible test specimen.<br />

• If yes then mark the test specimen <strong>for</strong> docking <strong>and</strong> retain the board number on the<br />

specimen with the letters “rb” <strong>for</strong> “r<strong>and</strong>om <strong>and</strong> biased”.<br />

R<strong>and</strong>om length measured from the end with the full board identification code<br />

End with the full board identification code<br />

Maximum strength reducing defect<br />

• If no then locate the maximum strength reducing defect in the board <strong>and</strong> position the<br />

test specimen template so that the defect is in the middle third of the test span.<br />

• Mark the test specimen ends <strong>for</strong> docking <strong>and</strong> retain the board number on the test<br />

specimen with the letters “b” after the sample number (“b” indicates a biased sample).<br />

• If the r<strong>and</strong>om sample doesn’t overlap the biased sample, two samples can be cut.<br />

84


R<strong>and</strong>om length measured from the end with the full board identification code<br />

End with the full board identification code<br />

Maximum strength reducing defect<br />

• If the biased <strong>and</strong> r<strong>and</strong>om samples overlap slightly, the combined length is docked <strong>and</strong><br />

the biased test section tested first, followed by the r<strong>and</strong>om section length.<br />

R<strong>and</strong>om length measured from the end with the full board identification code<br />

End with the full board identification code<br />

Maximum strength reducing defect<br />

• If the overlap is too large only one test can be done the piece with the priority has to<br />

be selected (r<strong>and</strong>om or biased). In this case only one specimen will result.<br />

R<strong>and</strong>om length measured from the end with the full board identification code<br />

End with the full board identification code<br />

Maximum strength reducing defect<br />

This protocol would ensure that <strong>for</strong> approximately 60% of boards, both r<strong>and</strong>om <strong>and</strong> biased<br />

test specimens will be available.<br />

85

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!